You are on page 1of 17

Journal of Membrane Science 232 (2004) 45–61

Analysis of modified surface force pore flow model with concentration


polarization and comparison with Spiegler–Kedem
model in reverse osmosis systems
Semant Jain1 , Sharad K. Gupta∗
Department of Chemical Engineering, Indian Institute of Technology, New Delhi 110016, India

Received 3 July 2003; received in revised form 4 November 2003; accepted 24 November 2003

Abstract

The Modified Surface Force Pore Flow (MD-SF-PF) model is used for predicting the performance of the sodium chloride–water and
sodium sulphate–water reverse osmosis systems. Unlike the previous analysis available in literature on this model, the present work takes
concentration polarization into account explicitly. The required mass transfer coefficient is estimated by relating it to the feed flow rate through
two additional parameters. The model equations being non-linear are solved using the orthogonal collocation technique. The model solution
code is validated by comparing results with those available in literature. Simulation studies indicate observed rejection versus flux has a
similar trend to that predicted by the Spiegler–Kedem model. From some of the experimental observations of Murthy [Studies on membrane
transport models, Ph.D. thesis, I.I.T. Delhi, July 1996], model solution and mass transfer parameters are estimated through a combination of
the Downhill Simplex method and Monte Carlo search. Simulated results using above parameters are found to be in very good agreement
with the remaining experimental observations. Prediction of the membrane performance and mass transfer coefficient are also found to be
very close to those estimated through the Spiegler–Kedem model. The membrane specific parameters such as the membrane thickness and
pore radius are calculated from experimental data for the two different systems on a similar cellulose acetate membrane. The closeness of
these parameter values again shows the validity of MD-SF-PF model.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Concentration polarization; Downhill simplex; Orthogonal collocation; Sodium chloride; Sodium sulphate; Spiegler–Kedem

1. Introduction applied across the membrane, against the direction of the


osmotic pressure gradient, then this tendency is reversed;
When a semi-permeable membrane (permeable to sol- i.e., the solution becomes more concentrated while the
vent, impermeable to solute) is placed between two solvent in it passes to the solvent rich side.
compartments—one containing pure solvent and the other Reverse Osmosis has become a very popular method for
containing solution—due to high osmotic pressure in the water purification. The major advantage of the process is
solvent side, solvent tends to go to the solution side. But its ability to perform direct separation at ambient conditions
if a pressure gradient greater than the osmotic pressure is and allowing separation of heat sensitive materials. It is cur-
rently being used extensively in the production of potable
and industrial water, treatment of domestic and industrial
Abbreviations: BC, back calculated; CFSK, combined film wastewater, and in food processing.
theory—Spiegler–Kedem model; E, Expt., experimental observation; Hr.,
hour; IT, irreversible thermodynamics; Lt., litre; P, value as published in
Development of a reliable membrane transport model for
literature; rms, root mean square; S, Sim., MD-SF-PF model simulated describing mass transport processes in a reverse osmosis
result; SK, Spiegler–Kedem model membrane is a highly desirable goal since not only does
∗ Corresponding author. Tel.: +91-11-26591023;
this allow the analysis of performance characteristics of a
fax: +91-11-26581120. membrane, but this can also assist in the design of new
E-mail addresses: semant@umich.edu (S. Jain),
sgupta@chemical.iitd.ernet.in (S.K. Gupta).
membranes. Onsager [1,2] used the principles of irreversible
1 Present address: Department of Chemical Engineering, University of thermodynamics to relate the fluxes with the forces through
Michigan, Ann Arbor, MI 48109, USA. phenomenological coefficients. However, a major limitation

0376-7388/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2003.11.021
46 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

of irreversible thermodynamics is that it is valid for sys- where the observed rejection is maximum, is predicted more
tems not far from equilibrium. The next step was the de- accurately by the CFSK model as compared to other mod-
velopment of Kedem–Katchalsky model [3–5] for a dilute els [13,14]. Thus, the CFSK model has been selected for
two-component system, consisting of water and a solute. comparing the predictions of our work.
The three adjustable parameters in this model are simple The Surface Force Pore Flow model [15] is a relatively
functions of the original phenomenological coefficients. In recent two-dimensional model for describing the transport
the Spiegler–Kedem model [6], the assumption on the ap- processes in a reverse osmosis membrane. Since the model
plicability of the linear laws throughout the thickness of the takes into account the membrane structure and various
membrane was resolved by rewriting the original linear IT membrane–solute interactions, it is expected the model will
equations in the differential form and then integrating them present a more accurate description of the mass transfer
over the membrane thickness. Although the Spiegler–Kedem process taking place outside the membrane.
model [6] gives quite accurate results, it is not a mechanis- Mehdizadeh and Dickson [16] pointed out the SF-PF
tic based model; i.e., it does not explain the mechanism of model uses an incorrect form of the material balance, the
transport or the nature of the membrane structure. potential function in the pore is inconsistent with the cylin-
Out of the mechanistic based models, the Solution Diffu- drical geometry and the solute concentration just inside the
sion model [7,8] has been applied to many different systems pore is equal to that of the permeate. These corrections led
as it is simple and has only two adjustable parameters and it to the development of the Modified Surface Force Pore Flow
is only appropriate for systems where the rejection is close model. The model contains four parameters—namely, θ 1
to unity. and θ 2 , required for estimating the potential function, RW ,
The three parameter Solution Diffusion Imperfection the membrane pore radius, and τ/ε, the ratio of the mem-
model [9] allowed for the presence of some imperfections, brane tortuosity to its porosity. Due to the complexity of the
which permit the salt to pass undiluted, but dodged the dif- non-differential equations of the MD-SF-PF model, numer-
ficulty of the explicit inclusion of the convection or viscous ical techniques are required for obtaining the solution such
flow. as the orthogonal collocation scheme [17,18].
The four parameter Finely Porous model [8,10,11] is The paper in which the MD-SF-PF model was defined
based on a balance of applied and frictional forces in a [16] assumed the solute bulk concentration is equal to the
one-dimensional pore, while the Frictional model [5,6] has solute concentration on the membrane surface and used this
a very simple physical interpretation, which amounts to a concentration to compute the theoretically observed rejec-
heuristic derivation. Both these models keep all the mech- tion. Their later work [19] mentions the concentration po-
anisms in their equations from the start but Mason and larization model can also relate the bulk feed concentration
Lonsdale [12] felt the treatment of viscous flow was either to the membrane surface concentration. However, detailed
misinterpreted or mishandled in both of them. analysis of the same was not carried out in their reference
The Diffusion Viscous flow or the Highly Porous model [20]. As concentration polarization effects are very impor-
[11] uses the Poisuielle’s law to describe the total volume tant in reverse osmosis, they must be included for model
flux and the total solute flux is expressed as the sum of a prediction and parameter estimation analysis.
contribution due to viscous and diffusive flow. In order to incorporate the concentration polarization
Mason and Lonsdale [12] reviewed all membrane trans- model into the MD-SF-PF model, the value of the mass
port models described above and compared them with the transfer coefficient is required. The mass transfer coefficient
statistical mechanical theory of membrane transport. They is dependent on several factors like cell geometry, feed flow
concluded all these models are equivalent in the sense that rate and temperature. Although its value can be computed
they end up writing down the same transport equations. for fully developed flow in well-established geometries like
However, these models differ in their predictions about channels and pipes, in actual experimental set-up for reverse
the transport coefficients. They finally recommended the osmosis the exact cell geometry and flow regime are often
Spiegler–Kedem model for describing and predicting the unknown. This introduces the need to relate the value of
performance of reverse osmosis and ultra filtration mem- the mass transfer coefficient to the flow rate for a specified
branes. solute–solvent system.
As the Spiegler–Kedem model relates the membrane sur- The objective of this work is to incorporate concentration
face concentration to the permeate concentration, it needs to polarization in the MD-SF-PF model and to develop a nu-
be combined with concentration polarization if the permeate merical solution for solving the differential equations of the
concentration is to be related to the bulk feed concentration model. The mass transfer coefficient may be related to flow
which results in the combined film theory—Spiegler–Kedem rate through the use of two additional parameters—‘a’ and
or CFSK model. Murthy [13] also analyzed the predictions ‘b’—increasing the number of parameters to be estimated
of the many membrane transport models with his experi- to six. Thus, a numerical scheme is developed for estimat-
mental observations and found the CFSK model to give the ing the parameters for a particular solute–solvent system. To
closest results. On plotting the reciprocal of the observed re- evaluate the accuracy of the MD-SF-PF model predictions,
jection against volume flux, he observed that the JVmin value, the estimated parameter set is used to predict the membrane
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 47

performance for a given system for the entire range of op- The Eqs. (1), (2) and (6) are the basic equations of the
erating conditions with the results being compared with the combined film—Spiegler–Kedem model (CFSK). By using
experimental observations. Additionally, the mass transfer a non-linear parameter estimation technique, the unknown
coefficient parameters and extrapolated trends of the recip- parameters, ‘PM ’, ‘σ’ and the mass transfer coefficient, ‘k’,
rocal rejection ratio to the volume flux have been compared can be determined from the given experimental data of ob-
with predictions of the CFSK model. served rejection ratio versus volume flux.
When the observed rejection ratio is plotted against vol-
ume flux for reverse osmosis systems, Murthy [13] found
2. Theory from his experimental data that a maximum in the observed
rejection is observed. The volume flux, JVmin , where this
2.1. Membrane mass transport models maximum occurs can be found from Eq. (6) by differenti-
ating it with respect to JV and then setting it to zero. The
The general purpose of a membrane mass transport model following relations are obtained:
is to relate the performance (usually expressed in terms of ln(1 + Pe∗ )
fluxes of solute and solvent and observed rejection) to the JVmin = k (7)
Pe∗
operating conditions (usually expressed in terms of pressure
and concentration driving forces). In a model, some coeffi- where
cients emerge that must be determined based on some ex- (1 − σ)k
Pe∗ = (8)
perimental data. The success of a model can be measured PM
in terms of its ability to describe mathematically the data
with coefficients that are reasonably constant over the range In the above relations, JVmin is the value of JV where a min-
of operating conditions. Here, we briefly discuss two mod- imum of (1/RO − 1) or a maximum in observed rejection
els: IT based Spiegler–Kedem model and mechanism based would occur at the value of JV given by Eq. (7).
Modified-Surface force-pore flow model which are used for The initially decreasing and then increasing trend of the
comparison in this study. reciprocal of observed rejection, (1/RO − 1), can be ex-
plained as follows. At low volume fluxes, the second term
2.2. Spiegler–Kedem model on the right in the last form of Eq. (6) can be neglected. As
the other terms in the equation are not affected by pressure,
The Spiegler–Kedem [6] model predicts the following (1/RO − 1), decreases with an increase in JV . On the other
equations for volume flux, JV , and actual rejection, RA : hand, at high volume fluxes, as the contribution of JV to
the first term becomes smaller, (1/RO − 1), is expected to
JV = LP (p − σ π) (1) increase exponentially with JV .
ePe − 1
RA = σ (2) 2.3. Modified-surface force-pore flow model
ePe − σ
where This work is based on the equations proposed in the
 
JV (1 − σ) MD-SF-PF model [16]. A few key equations are repro-
Pe = exp (3)
PM duced below while their derivation can be found from the
above-mentioned reference.
and ‘LP ’ is the hydraulic permeability coefficient of the
The differential equation for the velocity profile inside a
membrane; ‘PM ’ is the overall permeability coefficient and
pore can be found by a force balance in the z-direction on a
‘σ’ is the reflection coefficient.
fluid element in the annular region between z and z + dz and
On combining the relationship between RA , the actual re-
between r and r + dr inside a pore. The detailed derivation
jection, and RO , the observed rejection, with the film theory,
takes into account the net force due to difference in pressure,
the following equations are obtained:
  viscous shear stresses using Newton’s law of viscosity and
CA2 − CA3 JV the net force due to friction between the solute and the pore
= exp (4)
CA1 − CA3 k wall, leading to the following equation:
Eliminating CA1 , CA2 , and CA3 in terms of the actual rejec- d2 α(ρ) 1 dα(ρ) P − Π(1 − e−Φ(ρ) )
tion, RA , and observed rejection, RO , we obtain + +
  dρ2 ρ dρ β1
1 − RO 1 − RA JV −Φ(ρ)   
= exp (5) α(ρ) e 1 Π
RO RA k − 1− 1 + α(ρ) =0 (9)
β1 b(ρ) e −1
    
1 − RO 1−σ 1 JV where
ln = ln +
RO σ 1 − exp(−Pe) k ηDAB
β1 = (10)
(6) R2W π2
48 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

p wall. After integrating solute and solvent fluxes over the area
P = (11)
π2 of a single pore, they have been generalized over the sur-
face area of the membrane. In order to relate average solute
π2 − π3
Π = (12) and solvent fluxes through a single pore to the flux through
π2 a membrane, the fractional pore area on the membrane face,
r ε, is used. Thus, one gets:
ρ= (13)
RW
NA = ε J̄A (21)
The boundary conditions for Eq. (9) are:
NB = ε J̄B (22)
dα(ρ)
=0 at ρ = 0 (14)
dρ Then, the total flux is given by:
α(ρ) = 0 at ρ = 1 (15) NT = NA + NB (23)

The potential function used by SF-PF model had the poten- The permeate concentration, CA3 , is related to the solute and
tial function coming to a peak at the pore centerline instead solvent fluxes as:
of the expected smooth function. In the MD-SF-PF model J̄A
[16], it was proposed that an empirical equation could be ap- CA3 = C (24)
J̄A + J̄B
plied to any solute (electrolyte or non-electrolyte) and mem-
brane (charged or uncharged): The analysis in this paper assumes all pores on the membrane
 surface have the same radius. However, in case a pore size
 θ1
exp(θ2 ρ2 ), when ρ < 1 − λ distribution is present, then the solute and solvent fluxes
Φ(ρ) = RW (16)
 very large and positive, when ρ ≥ 1 − λ need to be integrated across all the possible pore sizes.

where 2.4. Incorporation of concentration polarization model in


RS the MD-SF-PF model
λ= (17)
RW
During the membrane separation process when the mem-
In the above equations, ‘θ 1 ’ affects the centerline potential brane rejects the solute, the solute concentration near the
function and ‘θ 2 ’ affects the radial slope of the potential. The membrane surface increases. Hence, there is a difference
radial gradient at the pore centerline is zero for this function. in the feed concentration in the bulk, CA1 , and that at the
An expression is required in the SF-PF or the MD-SF-PF membrane face, CA2 . The MD-SF-PF model assumes CA2 is
model, which describes the friction between the solute and known [16,19]. But since it is not possible to measure CA2 ,
the pore wall. This friction is the result of the hydrodynamic researchers have used the concentration polarization model
drag for a solute molecule moving in a small pore. It was first to relate it to CA1 . The build up of solute concentration in
suggested to use an empirical relationship for the friction the boundary layer region can be described by the film the-
function, ‘b’ [15]. In the later papers [21,22] a modified form ory. Thus, on rewriting the first form of Eq. (8) in terms of
of the Faxen equation has been used: concentrations, one obtains:
  
 1 NA + NB
, when λ ≤ 0.22 CA2 = CA3 + (CA1 − CA3 ) exp (25)
b(ρ) = 1 − 2.104λ + 2.09λ − 0.95λ
3 5
kC

44.57 − 416.2λ + 934.9λ + 302.4λ , when λ > 0.22
2 3
Although the concentration polarization can be easily inte-
(18) grated with the MD-SF-PF model, it has not been done so far.
Although the friction function should be a function of the After its addition, the computations become slightly more
radial position, the modified Faxen equation is a reasonable involved on account of two convergence loops that are now
approximation [16,19] and has been used in our analysis. needed [23]. The algorithm describing the key steps for in-
The solute and solvent fluxes have been averaged over the corporating the concentration polarization in the MD-SF-PF
pore cross-sectional area to obtain: model has been described in Jain [23].
  
2 1 1−λ α(ρ) π 2 − π3 2.5. Mass transfer coefficient
J̄A = π2 + exp(−Φ(ρ))
XAB τ 0 b(ρ) exp(α(ρ))−1
(19) The film mass transfer coefficient has been found to be
 1 dependent on feed flow rate, cell geometry, temperature and
2 1 solute system. A generalized correlation of mass transfer
J̄B = CRT α(ρ)ρ dρ (20)
XAB τ 0 suggests that the Sherwood number is related to the Reynolds
and Schmidt numbers as [24]:
In the above integration, the solute molecule has been as-
sumed to be no closer than one solute radius from the pore Sh = x1 Rex2 Scx3 (26)
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 49

Table 1
Physical properties estimation
S. No. Property Units Solute-1 Solute-2

1. Solute name NaCl Na2 SO4


2. Cationic conductance mho-m2 /equivalent 0.0050 0.0050
3. Anionic conductance mho-m2 /equivalent 0.0076 0.0080
4. Electrolyte conductance mho-m2 /equivalent 0.0126 0.0130
5. Cationic valence – 1 1
6. Anionic valence – 1 2
7. Diffusivity m2 /s 1.61 × 10−9 1.23 × 10−9
8. Solute radius m 1.3555 × 10−10 1.7743 × 10−10

where x1 , x2 and x3 are system and flow regime dependent to be diluted infinitely, equated to DAB for estimation of the
coefficients. solute radius. Values of the above parameters for NaCl–H2 O
In literature, Hwang and Kammermeyer [25] used the and Na2 SO4 –H2 O systems are shown in Table 1.
above equation for turbulent flows and Porter [26] in fully At 25 ◦ C and 1 atm, the density of water is taken as
developed and laminar regimes. All of them they took x3 = 1000 kg/m3 (1 g/cm3 ) and viscosity as 1 × 10−3 Pa s (1 cP)
0.33 which makes us believe that x3 can be taken as a con- [31]. These values are assumed to be constant throughout the
stant across flow regimes. With this in mind, Eq. (26) can operating conditions. Mehdizadeh and Dickson [27] used the
be simplified to obtain: value of viscosity of water as 8.965 × 10−4 Pa s (0.8965 cP).
k = a Qb (27) 2.7. Experimental observations
where ‘a’ and ‘b’ are the unknown parameters for a particular
solute–solvent combination. In order to estimate the parameters of any membrane
Murthy [13] has shown the value of ‘a’ depends on the transport model, experimental data are required. For the
solute–solvent system, while the value of ‘b’ ranges from present work, experimental data available in literature
0.33 to 0.8. [13,32] has been used. This section briefly mentions the
experimental conditions at which the above data was ob-
2.6. Physical properties estimation tained. More details of experimental set-up and conditions
can be obtained from the above references.
Some physical properties need to be estimated before the The experimental data was obtained in sets of inlet con-
MD-SF-PF model can be used for different solute–solvent centrations of 17.10, 103.18 and 207.62 mol/m3 (1000,
systems. These include estimating the solute radius and the 6000, 12,000 ppm) for NaCl–H2 O and 7.00, 42.51, 85.53,
diffusivity. While discussing the use of orthogonal colloca- 158.42 and 217.80 mol/m3 (1000, 6000, 12,000, 22,000
tion for solving the membrane equation [27] and propos- and 30,000 ppm) for Na2 SO4 –H2 O systems. For each inlet
ing the MD-SF-PF model [16], the values of diffusivity and concentration, flow rate is kept at 5.0 × 10−6 , 1.0 × 10−5 ,
solute radius for sodium chloride were specified. However, 1.5 × 10−5 , 2.0 × 10−5 and 2.5 × 10−5 m3 /s (300, 600,
the complete procedure used to obtain these values was 900, 1200 and 1500 ml/min). For each flow rate, pressure
not stated. Moreover, it was observed the DAB values for is kept at 2020, 3030, 4040, 6060, 8080 and 10,100 kPa
NaCl–H2 O in the above two papers also differed slightly. (20, 30, 40, 60, 80 and 100 atm). These three conditions,
In the present work, the solute molecular radius has been taken together, specify one set of operating conditions. For
estimated using the Stokes–Einstein equation [28]: a specific solute–solvent system, the concentration is ini-
tially kept constant. For each concentration, the flow rate
k0 T is gradually increased, and at each flow rate, pressure is
RS = (28)
6πηDAB incremented. A sample data set collected on experimental
observations on NaCl–H2 O system using Cellulose Acetate
where the value of diffusivity of strong electrolytes at infinite
membrane is shown in Table 2.
dilution may be calculated from an equation obtained by
Nernst on the assumption of complete dissociation [29]:
'0+ '0− z+ + z− 3. Solution of membrane transport equations
DA
0
= 8.931 × 10−10 T (29)
'0+ + '0− z+ z−
3.1. Orthogonal collocation method
A useful tabulation of ionic conductance at infinite dilution
in water at 25 ◦ C is given in Lange’s Handbook of Chemistry The SF-PF model used the Runge–Kutta technique to
[30], which has been used in the above equation. These solve the nonlinear differential equations [15]. This tech-
values are used to predict DA 0 and then, assuming the solution nique has been found to be slow, inefficient and expensive
50 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

Table 2 geometry) can be obtained from Table 3 of Villadsen and


Sample experimental data Stewart [17].
S. No. Pressure (Pa) CA3 (mol/m3 ) JV (m/s) RO

1. 2.02 × 106 4.0844 1.62 × 10−6 0.7613


3.1.3. Residual function
2. 3.03 × 106 2.8353 2.81 × 10−6 0.8343 For the system under consideration, the residual function
3. 4.04 × 106 2.1355 4.46 × 10−6 0.8752 at the ith collocation point becomes:
4. 6.06 × 106 1.6598 6.87 × 10−6 0.9030
× 106 × 10−6 n+1

5. 8.08 1.5075 9.44 0.9119 P − Π[1 − exp(−Φ(ρi ))]
6. 1.01 × 107 1.3980 1.28 × 10−5 0.9183 R|ρi = [Bij α(ρj )] +
β1
j=1
System: NaCl–H2 O; CA1 : 17.1111 mol/m3 ; Q: 5.0 × 10−6 m3 /s.
  
α(ρi )exp(−Φ(ρi )) 1 Π
for solving the SF-PF model equations [27]. The orthogonal − 1− 1+
β1 b exp(α(ρ)) −1
collocation of weighted residuals is used to provide an accu-
(35)
rate and efficient solution to the nonlinear differential equa-
tions in the SF-PF and MD-SF-PF models. A brief mention Theoretically, the value of the above function should be zero
of the relevant details is made below and further details can at all collocation points. However, as the values of the pa-
be obtained elsewhere [17,18]. rameters (a0 , a1 and a2 ) are not known a priori, the resid-
ual function will have some finite value. As the iterations
3.1.1. Interior collocation progress and the parameters approach their optimal value,
In order to estimate a dependent variable, in the present the residues tend to zero.
case—dimensionless velocity, it is approximated by a se-
ries of expansion containing n undetermined parameters. For 3.1.4. Updating guess
a symmetric second-order boundary value problem in one The Newton–Raphson technique is employed to update
variable—in the present case the dimensionless velocity—as the guess of the parameters used for the velocity profile:
a function of the dimensionless radial position, a suitable
F(xn )
form is: xn+1 = xn − (36)
J(xn )
n−1

y = y(1) + (1 − x2 ) ai Pi (x2 ) (30) In the above equation, ‘x’ represents the vector set of un-
i=0 known parameters (a0 , a1 and a2 ), ‘n’ the iteration number,
‘F’ the residual function value matrix and ‘J’ the Jacobian
In the present case, as α(1) = 0, assuming the validity of the
matrix.
no slip boundary condition, the above equation reduces to:
n−1
3.1.5. Termination condition
α = (1 − ρ ) 2
ai Pi (ρ2 ) (31) Iterations to obtain the optimal set of parameters are con-
i=0 tinued until the residues at all the collocation points are
within the tolerance limit of 10−10 and the maximum change
Using the values of the polynomials Pi (x2 ) as given in Table
in the value of any of the unknown parameters is also within
1 [17], for cylindrical geometry and n = 3; i.e. 3 unde-
the tolerance limit of 10−10 .
termined constants (a0 , a1 and a2 ), the velocity profile be-
comes:
3.2. Analytical solutions
α(ρ ) = (1 − ρ )[a0 + a1 (1 − 3 ρ )
2 2 2
Analytical solutions have been used to test the validity
+ a2 (1 − 8 ρ2 + 10 ρ4 )] (32)
of the MD-SF-PF model [16,27]. In the present work, we
too used the formulae under the Poiseuille flow and total
3.1.2. Interior formulae
rejection conditions to test our velocity profile trends under
The gradient and laplacian operators, in cylindrical coor-
identical operating conditions as specified in the above pa-
dinates, for the α(ρ2 ) of Eq. (32) are given by:
pers and found that our estimates matched those published

n+1

dα(ρ2 )

excellently [23].
= Aij α(ρj2 ) (33)

ρ=ρi
j=1

 
n+1
4. Parameter estimation
1 d 1 dα(ρ2 )



= Bij α(ρj2 ) (34)
ρ dρ ρ dρ ρ=ρi The modified surface force pore flow model along with the
j=1
concentration polarization model contains six parameters—
for i = 1, 2, . . . , n + 1. The coefficients Aij and Bij which θ 1 , θ 2 , τ/ε, RW , a and b. These parameters are determined
have been used in this work (for symmetrical cylindrical by the Downhill Simplex method as described below.
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 51

Table 3
Data analysis for NaCl–H2 O system
S. No. Parameter Unit Value

(A) Parameters used for data analysisa


1. θ1 m 1.5007 × 10−9
2. θ2 – 4.8739
3. RW m 9.9917 × 10−10
4. τ/ε m 1.2644 × 10−4
5. b – 0.7786
6. a (m/s)/(mol/m3 )0.7786 1.0733
(B) Data analysis using above parameters
(1) Operating condition
1.1. CA1 (mol/m3 ) 17.11 103.18 207.62
1.2. Conditionb BC BC P
(2) Error analysis (% rms averaged)c
2.1. CA3 6.17 8.63 10.27
2.2. JV 17.39 21.76 21.71
2.3. RO 1.50 2.35 2.57
2.4. k 3.15 2.18 2.05
2.5. E 13.05 16.55 16.98
a Note: these parameters are averaged parameters obtained from observations at 17.11 and 103.18 mol/m3 .
b ‘BC’ means ‘back calculated’, implying that the experimental data set is used for parameter estimation. ‘P’ means ‘predicted’, implying that the data
set has not been used in parameter estimation in any way.
c Error analysis compares experimental observations to and simulation estimations for C , J and R . k values are compared with CFSK model
A3 V O
predictions. E is the overall objective error function.

4.1. Optimization problem 4.2. Downhill simplex method

Experimental values are available in terms of operating At this stage we have a six-parameter non-linear optimiza-
conditions (inlet concentration, pressure drop and flow rate), tion problem. We could not identify a technique that would
permeate solute concentration and flux. The model equations guarantee the global minimum solution to the problem, so
are solved from initial guesses of the six parameters and the we combined an approach giving a local minimum and de-
results compared with the experimental observations. Two veloped a search technique that would optimize the solution
optimization functions are formed, first one, E1 , from min- [23]. We selected the Downhill Simplex method [33]. As
imization of the fractional difference in the calculated and starting with any initial guess, this approach almost always
experimental permeate concentration and the other, E2 , from converges to a local minimum.
minimizing the fractional difference between the calculated The manner in which the optimization operations are car-
and experimental flux: ried out and estimation of intermediate values has been car-
ried out follow the development in literature [33,34].
n i i
i=1 (1 − (CA3 |Sim /CA3 |Expt ))
2
These following sub-sections briefly describe some of the
E1 = (37)
n important issues we had to address while using the Downhill
Simplex method. More details can be found in Jain [23].
n i i
i=1 (1 − (JV |Sim /JV |Expt ))
2
E2 = (38) 4.2.1. Initial guess
n
Ideally, an optimization technique should be independent
The fractional errors are used instead of absolute differences, of the initial guess. However, since the Downhill Simplex
as this would make the values independent of the units used, method is a local optimization technique, it is important to
thus preventing a bias towards any one parameter. On com- have a very good initial guess. At first, the order of mag-
bining these two functions, the overall objective function nitude of each of the parameters is determined (θ 1 : 10−9 ,
becomes: θ2 ∼ 5, RW : 10−9 , τ/ε: 10−4 , a ∼ 1.0 and b ∼ 0.7). This
is followed by simulation runs in which intermediate steps
E1 + E 2
E= (39) are analyzed to check for any physical inconsistency. In case
2 none is observed, then the set is selected for preliminary
The aim of the optimization procedure is to find out the simulation purposes. This initial guess is then updated as
values of θ 1 , θ 2 , RW , τ/ε, a and b at which E is very small, more simulations are carried out in an attempt to improve
i.e., the theoretical predictions match the experimental ones the overall error. The automated initial guesses updation al-
with minimum error. gorithm can be found in Jain [23].
52 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

4.2.2. Convergence criterion constant. In order to explore the nearby surface, with small
The convergence of the Downhill Simplex method de- variations in the values of the parameters, an approach simi-
pends on the values of arbitrary constants in the reflection, lar to that adopted in genetic algorithms has been developed
contraction and expansion operations. It has been observed along with the Downhill Simplex loop [23]. The algorithm
fast convergence is obtained by taking α = 1, β = 0.5 and of the Monte Carlo search technique has been described in
γ = 2 [33,34]; therefore, we adopted the same values for Appendix A.
the present analysis.
The convergence criterion is not applied with respect to
the changes in each optimization parameter but to the value 5. Results
of the total functions, E1 and E2 , represented as a whole in
E. The reasoning behind this approach is that in statistical 5.1. Validation
problems where one is concerned with finding the minimum
of a negative likelihood surface (or of a sum of a square sur- The first aspect of the project is to solve the differential
face) the curvature near the minimum gives the information equation for the velocity profile using the orthogonal collo-
available on the unknown parameters. cation method. The present model solution code is validated
The objective function is considered locally optimized by comparing results with those presented by Mehdizadeh
when its value becomes a constant while using the centroid and Dickson [27] under identical conditions. For this sec-
parameters for computing its value. It is observed its value tion, the same values of physical properties; i.e., diffusivity,
comes within 0.2 well before 30 iterations of the Downhill solute radius and viscosity are taken which were used in the
Simplex loop. Further iterations are not carried out since original work [27]. Since, in the present code, the concen-
no change is observed in the parameters and the error is a tration polarization is taken into account while Mehdizadeh
constant. The above approach is acceptable since using the and Dickson [27] had neglected it, CA2 is forced to be nearly
Downhill Simplex method often leads to local minima. equal to CA1 by making mass transfer coefficient quite large.
This is achieved by fixing ‘a’ at 1000, about 108 times the
4.2.3. Interdependent parameters value normally observed when predicting the mass trans-
Another issue to be considered is that all the five parame- fer coefficient from experimental data, and ‘b’ as 0, which
ters affect both the error functions—E1 and E2 . In case either makes the mass transfer coefficient independent of flow rate
flux or permeate concentration is to be changed, selecting and its value sufficiently high for it to be considered infinite.
any one parameter will affect both of them. For example, if Our computed velocity profiles, actual rejection ratios and
flux was to be decreased, one could think of increasing the flux ratios are found to be in excellent agreement [23] with
potential function, meaning increasing θ 1 and/or θ 2 . How- those published by Mehdizadeh and Dickson [27] validating
ever, as soon as the initial guess of the parameter is changed, the present orthogonal collocation code.
in the orthogonal collocation convergence, the coefficients
a0 , a1 and a2 will change. This would imply change in α(ρ), 5.2. Concentration polarization model
thus affecting JB and ultimately CA3 . Thus, it is very difficult
to predict the actual affect of changing a selected parameter. The effect of introducing concentration polarization to re-
late CA1 and CA2 is seen in Figs. 1 and 2. Parameters used
4.2.4. Space constraints are θ 1 : 5.4434 × 10−9 m, θ 2 : 0.4910, τ/ε: 1.0 × 10−4 m, RW :
In addition to the convergence criterion, there are con- 1.3 × 10−9 m, a: 1000 m/s and b: 0. The operating condi-
straints on the operational space as well; i.e., all the pa- tions include temperature at 298 K, pressure as 4000 kPa and
rameters are non-negative. This is incorporated by writing the bulk feed concentration as 75 mol/m3 . For these simula-
each parameter in the objective function as the square of its tion runs, viscosity of water is taken as 0.001 kg/ms (1 cP),
root. It is also ensured the parameters have physical corre- diffusivity as 1.566 × 10−9 m/s and solute (NaCl) radius as
lation throughout the simplex operations. A simple check 1.555 × 10−10 m (1.555 Å).
is maintaining the values of key physical quantities—CA2 , It is desired to represent a range of values for the mass
CA3 , NA , NB , NT and JV as non-negative. Additionally, the transfer coefficient for all flow regimes—turbulent, laminar
lower limit of CA2 is fixed at CA1 , upper limit of CA2 as and intermediate. This is achieved by maintaining parameter
twice CA1 , upper limit of CA3 as CA1 , and in accordance ‘b’ as 0, making mass transfer coefficient equal to ‘a’. The
with experimental observations, value of ‘b’ was maintained value of ‘a’ is changed from 103 to 5.0 × 10−6 m/s which
in the range of 0.33–0.8. covered the above flow regimes. Figs. 1 and 2 depict the
effect of increasing bulk feed concentration, with the mass
4.3. Monte Carlo search transfer coefficient as a parameter, on rejection ratio and
volume flux respectively. The decreasing trends of rejection
It is often observed that the Downhill Simplex method ratio and volume flux with increasing feed concentration are
attained local minima after a few iterations itself. As the pa- along the expected lines. On the other hand, at a particu-
rameters do not vary, the objective function ‘E’ becomes a lar value of the bulk feed concentration, both rejection ra-
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 53

Fig. 1. Effect of bulk feed concentration on observed rejection ratio with the mass transfer coefficient as a parameter. Operating conditions: T = 298 K,
P = 4000 kPa and CA1 = 75 mol/m3 for NaCl–H2 O system. Model parameters: θ1 = 5.4434 × 10−9 m, θ2 = 0.4910, τ/ε = 1.0 × 10−4 m and
RW = 1.3 × 10−9 m.

tio and volume flux increase with increasing mass transfer ization has already become insignificant and lower values
coefficient, which obviously is due to the reduced concen- than 5 × 10−6 m/s are not present in real reverse osmosis
tration polarization at higher mass transfer coefficients. The operations.
above graphs reflect the effect of mass transfer coefficient When using the Spiegler–Kedem model [6], on increasing
from 5 × 10−6 to 10−1 m/s. For higher values of the mass JV , 1/RO −1 initially decreases, attains a minimum, and then
transfer coefficients than 10−1 m/s the concentration polar- starts increasing (Section 2.2). To test this trend, the model

Fig. 2. Effect of bulk feed concentration on flux with the mass transfer coefficient as a parameter. Operating conditions: T = 298 K, P = 4000 kPa and
CA1 = 75 mol/m3 for NaCl–H2 O system. Model parameters: θ1 = 5.4434 × 10−9 m, θ2 = 0.4910, τ/ε = 1.0 × 10−4 m and RW = 1.3 × 10−9 m.
54 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

Fig. 3. Effect of concentration polarization on reciprocal of rejection ratio vs. flux with the mass transfer coefficient as a parameter. Operating conditions:
T = 298 K and CA1 = 75 mol/m3 for NaCl–H2 O system. Pressure is incremented in steps of 1000 kPa from a minimum of 1000 kPa to a maximum of
4000 kPa. Model parameters: θ1 = 5.4434 × 10−9 m, θ2 = 0.4910, τ/ε = 1.0 × 10−4 m and RW = 1.3 × 10−9 m.

solution code is run at a temperature of 298 K for a bulk feed function encompassing the effects on both observed rejec-
concentration of 75 mol/m3 of the NaCl–H2 O solution with tion ratio and the volume flux. This approach is followed
pressure being incremented in steps of 1000 kPa from 1000 as parameters like θ 1 and θ 2 must be examined together,
to 4000 kPa. The model parameters used are: θ 1 : 5.4434 × and not individually, for potential function evaluation. As
10−9 m, θ 2 : 0.4910, τ/ε: 1.0×10−4 m, RW : 1.3×10−9 m with the optimization process is fairly random, while θ 1 may be
the mass transfer coefficient having been varied from 5.0 × affected in some iterations, others times θ 2 may be changed.
10−6 to 103 m/s. Observations at mass transfer coefficient of With just 2 or 3 data sets being used for computing aver-
greater than 2.5×10−5 m/s show no discernible change from aged parameter sets, we felt carrying out an error analysis
the trend at 2.5 × 10−5 m/s and so have not been included would make more sense.
in the graph (Fig. 3). For the case of NaCl–H2 O system, all experimental data
Looking at the above graphs, it can be said the concentra- at 17.11 and 103.18 mol/m3 have been used separately for
tion polarization becomes significant when the mass trans- parameter estimation. When these parameters are used for
fer coefficient lies in 10−6 –10−4 m/s range—essentially the analysis of the same set of experimental data, the back
regime in which most of the experimental data sets lie. calculation errors (E) are 13.05 and 16.55% for 17.11 and
103.18 mol/m3 , respectively. The averaged parameters for
5.3. Model evaluation these two concentrations are given in Table 3. As shown
in this table, using these averaged parameters to estimate
Parameter estimation has been done in two steps. In the the membrane performance at 207.62 mol/m3 , the results
first step, all the experimental observations at some selected are found to be in excellent agreement with the experi-
bulk feed concentrations are used to estimate the parame- mental observations with ‘E’ being 16.98%. Subsequently,
ters. These parameters are then used to back calculate the these parameters are used to further predict the membrane
experimental observations from which they are produced in performance and the detailed results are discussed below
order to study the consistency of the original experimental (Section 5.4).
observation. When a similar procedure is repeated for Na2 SO4 –H2 O
In the second step, the above estimated parameters are system, with the parameters being estimated at 42.51 and
averaged to obtain the ‘average parameters’. These param- 158.42 mol/m3 , the back calculation values errors (E) are
eters are then used to predict the membrane performance of 15.51 and 14.24%, respectively. The averaged parameters for
all available experimental conditions including those, which these two concentrations are shown in Table 4. Using these
were not used in any way for parameter estimation. parameters, the errors (E) between the predicted and the ex-
For data analysis, we have preferred to examine error perimental observations for concentrations 7.00, 85.53 and
functions ‘E1 ’, ‘E2 ’ and ‘E’, rather than the parameters 217.80 mol/m3 are estimated and found to be 18.74, 12.90
individually. ‘E’ has been selected as the single point error and 16.98% respectively. The above values in ‘E’ clearly in-
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 55

Table 4
Data analysis for Na2 SO4 –H2 O system
S. no. Parameter Unit Value

(A) Parameters used for data analysisa


1. θ1 m 2.6144 × 10−9
2. θ2 – 7.3283
3. RW m 1.2090 × 10−9
4. τ/ε m 1.0404 × 10−4
5. b – 0.7769
6. a (m/s)/(mol/m3 )0.7769 0.8620
(B) Data analysis using above parameters
(1) Operating condition
1.1. CA1 (mol/m3 ) 42.51 158.42 7.00 85.53 217.80
1.2. Conditionb BC BC P P P
(2) Error analysis (% rms averaged)c
2.1. CA3 10.69 10.60 13.60 8.27 15.33
2.2. JV 19.15 17.13 22.75 16.26 18.49
2.3. RO 0.47 0.64 0.53 0.44 1.34
2.4. k 2.00 8.27 5.81 5.81 3.85
2.5. E 15.51 14.24 18.74 12.90 16.98
a
Note: these parameters are averaged parameters obtained from observations at 42.51 and 158.42 mol/m3 .
b
‘BC’ means ‘back calculated’, implying that the experimental data set is used for parameter estimation. ‘P’ means ‘predicted’, implying that the data
set has not been used in parameter estimation in any way.
c Error analysis compares experimental observations to and simulation estimations for C , J and R . k values are compared with CFSK model
A3 V O
predictions. E is the overall objective error function.

dicate the parameter estimation program has estimated mem- with flow rate as a parameter (Figs. 4 and 5) and the volume
brane parameters, which predict the membrane performance flux and permeate concentration against bulk feed concen-
for unknown conditions quite accurately. tration with flow rate as a parameter (Figs. 6 and 7).
Since the experiments on both the systems are conducted From Fig. 4, it can be seen the predictions match the ex-
on similar Cellulose–Acetate membrane, the membrane spe- perimental observations very closely. However, the rejection
cific parameters, RW and τ/ε, are expected to remain largely ratios are predicted slightly better than the fluxes. In Fig. 5,
constant. Their values and the error analysis of these param- the simulation results excellently match the experimental ob-
eters are compared for both the systems in Table 5. While servations. Slightly higher errors of rejection ratio at lower
making these membranes, Murthy [13] made several batches pressures are possible because of a lower than expected ex-
of membranes. Although the same preparation technique perimental observation of rejection ratio at 3000 kPa.
and conditions were kept while making these membranes, Figs. 6 and 7 have been prepared in order to show the
some small differences between batches may have remained. membrane performance across the entire range of exper-
This may be the reason for differences of 17.36% for RW imental conditions. For both the systems, the trend of
and 21.53% for τ/ε when the estimations from NaCl–H2 O permeates concentration against bulk feed concentration
and Na2 SO4 –H2 O systems were compared as shown in matches very closely with the experimental observations.
Table 5. For NaCl–H2 O system, the maximum error at a flow rate
of 5.0 × 10−6 m3 /s and concentration of 207.62 mol/m3 is
5.4. Membrane performance prediction observed to be 4.5%. The corresponding figure for flux at
the same operating conditions is observed to be 8.9%.
In addition to the objective error function ‘E’, the mem- In Fig. 7, for Na2 SO4 –H2 O system, the permeate con-
brane performance predictions are also carried out for the centration is plotted against bulk feed concentrations at a
volume flux and permeate concentration against pressure flow rate of 2.0 × 10−5 m3 /s. Errors between the MD-SF-PF
model predictions and the experimental observations in per-
meate concentration are just 2.6, 4.0 and 9.1% for bulk feed
Table 5
concentrations of 7.00, 85.53 and 217.80 mol/m3 , respec-
Comparison of membrane specific parameters
tively. The corresponding values of errors for the flux at the
S. no. System RW (m) τ/ε (m) same operating conditions are 9.3, 0.9 and 22.5%, respec-
1. NaCl–H2 O 9.9917 × 10−10 1.2644 × 10−4 tively. Figs. 4–7 and the above analysis clearly show the
2. Na2 SO4 –H2 O 1.2090 × 10−9 1.0404 × 10−4 MD-SF-PF model has been successful in estimating the pa-
3. Error (%)a 17.36 −21.53 rameters and in predicting the membrane performance very
a %Error = 100[1 − parameter(NaCl)/parameter(Na2 SO4 )]. closely.
56 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

Fig. 4. Effect of pressure on rejection ratio and flux for NaCl–H2 O system with flow rate (m3 /s) as a parameter. ‘S’ represents simulated results
and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K and CA1 = 207.62 mol/m3 (1000 ppm). Model parameters:
θ1 = 1.5007 × 10−9 m, θ2 = 4.8739, τ/ε = 1.2644 × 10−4 m and RW = 9.9917 × 10−10 m, b = 0.7786 and a = 1.0733 (m/s)/(m3 /s)0.7786 .

5.5. Mass transfer coefficient the coefficient parameters from the CFSK model using
the Box-Kanemasu method [35]. When the mass transfer
Mass transfer coefficient and its parameters ‘a’ and coefficient parameters from the two different methods are
‘b’ are computed from the present parameters estimation compared for both NaCl–H2 O and Na2 SO4 –H2 O systems in
analysis and compared with the values obtained from the Table 6, the maximum error in any parameter, for any sys-
combined film theory—Spiegler–Kedem (CFSK) model. tem, is around 3% implying the predicted values from MD-
Murthy [13] estimated the mass transfer coefficient and SF-PF and the CFSK models are in excellent agreement.

Fig. 5. Effect of pressure on rejection ratio and flux for Na2 SO4 –H2 O system with flow rate (m3 /s) as a parameter. ‘S’ represents simulated results
and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K and CA1 = 85.53 mol/m3 (1000 ppm). Model parameters:
θ1 = 2.6144 × 10−9 m, θ2 = 7.3283, τ/ε = 1.0404 × 10−4 m, RW = 1.2090 × 10−9 m, b = 0.7769 and a = 0.8620 (m/s)/(m3 /s)0.7769 .
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 57

Fig. 6. Effect of bulk feed concentration on permeate concentration and flux for NaCl–H2 O system with flow rate (m3 /s) as a parameter. ‘S’ represents
simulated results and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K and P = 4,040 kPa. Model parameters:
θ1 = 1.5007 × 10−9 m, θ2 = 4.8739, τ/ε = 1.2644 × 10−4 m, RW = 9.9917 × 10−10 m, b = 0.7786 and a = 0.0733 (m/s)/(m3 /s)0.7786 .

5.6. Extrapolated results sure range of 2,000–40,000 kPa (19.8–396.0 atm). Plots for
the systems under consideration are shown in Figs. 8 and
When the reciprocal of the observed rejection is plotted 9. Our results indicate the JVmin would occur at 23,500 kPa
against volume flux for both NaCl–H2 O and Na2 SO4 –H2 O (232.67 atm) for NaCl–H2 O and at 11,500 kPa (113.86 atm)
systems under consideration, minima is not found within the for Na2 SO4 –H2 O.
experimental region. This necessitated the need to study the The value of JVmin estimated through the MD-SF-PF
system in an extrapolated region; i.e., beyond the experimen- model for NaCl–H2 O system is 2.41 × 10−5 m/s and for
tal region. As the experimental data are limited to 10,100 kPa the Na2 SO4 –H2 O system is 2.01 × 10−5 m/s. Correspond-
(100 atm), the simulations are carried out to cover the pres- ing values for these systems estimated using the CFSK

Fig. 7. Effect of bulk feed concentration on permeate concentration and flux for Na2 SO4 –H2 O system with flow rate (m3 /s) as a parameter. ‘S’ represents
simulated results and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K and P = 10,100 kPa. Model parameters:
θ1 = 2.6144 × 10−9 m, θ2 = 7.3283, τ/ε = 1.0404 × 10−4 m, RW = 1.2090 × 10−9 m, b = 0.7769 and a = 0.8620 (m/s)/(m3 /s)0.7769 .
58 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

Fig. 8. Effect of flux on rejection ratio for NaCl–H2 O system. ‘S’ represents simulated results using MD-SF-PF model, ‘CFSK’ are values obtained through
combined film—Spiegler–Kedem model and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K, CA1 = 103.18 mol/m3 ,
Q = 1.5 × 10−6 m3 /s and P = 2,000–40,000 kPa. MD-SF-PF model parameters: θ1 = 1.5007 × 10−9 m, θ2 = 4.8739, τ/ε = 1.2644 × 10−4 m,
RW = 9.9917 × 10−10 m, b = 0.7786 and a = 1.0733 (m/s)/(m3 /s)0.7786 . Spiegler–Kedem model parameters: σ = 0.9395, PM = 4.168 × 10−7 m/s,
b = 0.7763 and a = 1.0355 (m/s)/(m3 /s)0.7763 [13].

model are 2.29 × 10−5 and 2.19 × 10−5 m/s, respectively. ing at the above error calculations and the overall trends
Assuming the CFSK model estimated values as the base, displayed in the figures, the MD-SF-PF model predictions
this means an error of −5.36% for the NaCl–H2 O system of JVmin are in excellent agreement with the CFSK estima-
and an error of 8.08% for the Na2 SO4 –H2 O system. Look- tions.

Fig. 9. Effect of flux on rejection ratio for Na2 SO4 –H2 O system. ‘S’ represents simulated results using MD-SF-PF model, ‘CFSK’ are values obtained through
combined film—Spiegler–Kedem model and ‘E’ the experimental observations of Murthy [13]. Operating conditions: T = 298 K, CA1 = 103.18 mol/m3 ,
Q = 1.5 × 10−6 m3 /s and P = 2000–40,000 kPa. MD-SF-PF model parameters: θ1 = 2.6144 × 10−9 m, θ2 = 7.3283, τ/ε = 1.0404 × 10−4 m,
RW = 1.2090 × 10−9 m, b = 0.7769 and a = 0.8620 (m/s)/(m3 /s)0.7769 . Spiegler–Kedem model parameters: σ = 0.9982, PM = 1.508 × 10−7 m/s,
b = 0.7769 and a = 0.8395 (m/s)/(m3 /s)0.7769 [13].
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 59

Table 6
Comparison of mass transfer coefficient parameters Nomenclature
S. No. Property NaCl–H2 O Na2 SO4 –H2 O a unknown in film mass transfer equation
1. Parameters a b a b (Eq. (27)), (m/s)/(m3 /s)b
a a
2. Units – – a0 unknown parameter for velocity profile
3. Predictionb 1.0683 0.7786 0.8620 0.7769 (Eq. (32))
4. CFSKc 1.0355 0.7763 0.8395 0.7769 a1 unknown parameter for velocity profile
5. Error (%)d −3.16 −0.29 −2.68 0.00
(Eq. (32))
Units of ‘a’: (m/s)/(m3 /s)b
a a2 unknown parameter for velocity profile
b
Values from simulation based on the MD-SF-PF model in the present (Eq. (32))
work. ai unknown parameter for the estimation of the
c Values from the CFSK model parameters have been obtained from
velocity profile (Eq. (30))
Murthy [13].
d %Error = 100[1 − k(Sim.)/k(CFSK)]. Aij first derivative coefficient matrix for
orthogonal collocation (Eq. (33))
b unknown in film mass transfer equation
(Eq. (27))
6. Conclusions b(ρ) frictional parameter (Eq. (18))
Bij second derivative coefficient matrix for
The addition of concentration polarization to the MD-
orthogonal collocation (Eq. (34))
SF-PF model resulted in a six-parameter estimation prob- C
lem which has been solved by using Monte Carlo search molar density of solution (mol/m3 )
CAi solute concentration at ith position (mol/m3 )
and automated guess updation combined with the Down-
i
CA3
hill Simplex technique. The estimated parameters are used solute concentration in permeate solution
to predict the membrane performance over a range of op- during ith observation (mol/m3 )
erating conditions, and the simulated results compare very DA
0
solute diffusivity in infinitely diluted solution
well with the experimental observations. The membrane spe- (m2 /s)
cific parameters of NaCl–H2 O and Na2 SO4 –H2 O systems DAB solute diffusivity in solution (m2 /s)
on a similar cellulose acetate membrane have been shown E objective error function (Eq. (56))
to be very close to each other. The rejection ratio versus the E1 error in permeate concentration (Eq. (54))
flux trend and the predicted JVmin values for the system un- E2 error in flux (Eq. (55))
der consideration show agreement with the predictions of F (xn ) residual function value matrix (Eq. (36))
Spiegler–Kedem model. All other observations indicate the
J̄A radially averaged solute flux through a single
success of MD-SF-PF model as its predictions are found
pore (mol/m2 s)
to be in agreement with the Spiegler–Kedem model—the J̄B radially averaged solvent flux through a
model recommended by Mason and Lonsdale [12] for re-
single pore (mol/m2 s)
verse osmosis membranes. Being a mechanism based model, J (xn ) Jacobian matrix (Eq. (36))
the advantage of the MD-SF-PF model when combined
JV volume flux (m/s)
with concentration polarization over the Spiegler–Kedem
model is that its parameters can be related to the mem- JVi volume flux at ith observation (m/s)
brane structure, the solute–solvent and solute–membrane JVmin value of flux, JV , when 1/RO − 1 is
interactions. minimum when plotted against JV (m/s)
k film mass transfer coefficient (m/s)
kO Boltzmann’s constant (J/K)
Acknowledgements l+
0
cationic conductance at infinite dilution
(mho/equivalent)
l−
0
anionic conductance at infinite dilution
The authors would like to thank Mr. J.K. Jain, Sys-
tems Programmer, Department of Chemical Engineering, (mho/equivalent)
0 + l0
l+ − electrolyte conductance at infinite dilution
IIT-Delhi, Mr. Om Vir Singh from the Department of Chem-
(mho/equivalent)
ical Engineering and Mr. Deepak Garg from the Department
LP hydraulic permeability coefficient of the
of Computer Science & Engineering who assisted us with
membrane (m2 s/kg)
the coding aspects. Assistance and support extended by Dr. n number of experimental observations
Rajesh Khanna and Dr. Kamal K. Pant, faculty members in
Ni flux of component ‘i’ through membrane
the Chemical Engineering department, is also gratefully ac-
(mol/m3 )
knowledged. The authors would also like to thank the two
NTi total flux during ith observation (mol/s)
anonymous reviewers whose numerous suggestions helped
improve the quality of the paper significantly.
60 S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61

p pressure drop across the membrane (Pa) β parameter used in downhill simplex method
P ratio of pressure drop across the membrane to β1 parameter used in model solution (Eq. (9))
osmotic pressure at boundary layer (Eq. (11)) χij frictional constant between i and j (J s/m2 mol)
PA solute permeability—CFSK ε fractional pore area of membrane
phenomenological parameter (m2 /s)
PB water permeability—CFSK Φ(ρ) dimensionless potential function
phenomenological parameter (m3 s/kg) γ parameter used in Downhill Simplex method
Pe Peclet number η solution viscosity (Pa s)
Pe∗ value of Pe, for which 1/RO − 1 is minimum λ ratio of solute radius to the membrane pore
when plotted against JV (m/s) radius
Pi polynomial for evaluation of the πi osmotic pressure at location ‘i’ (Pa)
dimensionless velocity profile (Eq. (30)) π osmotic pressure across the membrane (Pa)
PM overall membrane permeability coefficient Π ratio of difference of osmotic pressure at
(m/s) boundary layer and permeate to the osmotic
Q flow rate (m3 /s) pressure at boundary
r radial coordinate (m) θ1 unknown in surface potential function (m)
R gas constant (J/mol K) θ2 unknown parameter in surface potential
RA actual rejection ratio, difference of ratio of function
membrane surface concentration to the ρ ratio of radial position to pore radius
permeate concentration from 1 ρi radial position of ith interior collocation point
RO observed rejection ratio (Eq. (33))
Re Reynolds number σ reflection coefficient
R|ρi residual function evaluated at ith τ average pore length taking tortuosity into
dimensionless radial position (Eq. (30)) account (m)
RS solute radius (Eq. (28)) (m)
Subscripts
RW membrane pore radius (m) A solute
Sc Schmidt number B solvent
Sh Sherwood number E experimental observation
T temperature of solution (K) S simulated result
x independent variable for orthogonal T
collocation equations total solution
x1 experimentally determined coefficient for W pore wall
estimating Sh (Eq. (26)) 1 feed solution
x2 experimentally determined coefficient for 2 boundary layer solution
estimating Sh (Eq. (26)) 3 permeate solution
x3 experimentally determined coefficient for
estimating Sh (Eq. (26))
XAB ratio of the product of gas constant and
solution temperature to solute diffusivity in Appendix A. Monte Carlo search algorithm
solution (kg/mol s)
xn The Downhill Simplex method converges to a minimum,
vector set of parameters (a0 , a1 and a2 ) at nth
which is most likely, a local one. In an attempt to identify the
iteration (Eq. (36))
xn+1 vector set of parameters (a0 , a1 and a2 ) at global minima of the problem, the nearby surface solution
(n + 1)th iteration (Eq. (36)) needs to be explored. This is accomplished by adding a
x membrane thickness (m) Monte Carlo search algorithm [23]. The integration of the
y Monte Carlo search with the Downhill Simplex routine has
dependent variable for orthogonal collocation
been done in the following manner:
equations
z+ absolute value of cation valence 1. This procedure is activated at every third iteration of the
z− absolute value of anion valence simplex loop to start after four iterations of the simplex
loop have been completed. The objective of this step is
Greek letters to examine the nearby area of a potential local minima
α parameter used in downhill simplex method already identified by the downhill simplex method.
α(ρ) solute velocity 2. Values of ‘E’ for the past three iterations are compared.
If error becomes constant, or consistently increases, it in-
S. Jain, S.K. Gupta / Journal of Membrane Science 232 (2004) 45–61 61

dicates possible local minima might have been attained. [13] Z.V.P. Murthy, Studies on membrane transport models, Ph.D. thesis,
For all other cases the Downhill Simplex routine is per- I.I.T. Delhi, July 1996.
[14] Z.V.P. Murthy, S.K. Gupta, Sodium cyanide separation and param-
mitted to continue without any deviations. eter estimation for reverse osmosis thin film composite polyamide
3. Only the first six simplex vertices of the total seven are membrane, J. Membr. Sci. 154 (1999) 89–103.
examined as a six-parameter combination is being han- [15] T. Matsuura, S. Sourirajan, Reverse osmosis transport through cap-
dled. This is done to prevent any undue weightage being illary pores under the influence of surface forces, Ind. Eng. Chem.
assigned to any parameter. In case the parameter serial 20 (2) (1981) 273–282.
[16] H. Mehdizadeh, J.M. Dickson, Theoretical modification of the Sur-
number is equal to the vertex serial number, it is further face Force Pore Flow model for reverse osmosis transport, J. Membr.
tested (serial order of parameters: θ 1 , θ 2 , RW , τ/ε, a and Sci. 42 (1989) 119–145.
b). [17] J.V. Villadsen, W.E. Stewart, Solution of boundary value problems
4. The variation probability of a parameter’s value has been by orthogonal collocation, Chem. Eng. Sci. 22 (1967) 1483–1501.
fixed at 0.5. Thus, if the randomly generated probability [18] J.V. Villadsen, M.L. Michelsen, Solution of Differential Equation
Models by Polynomial Approximation, Prentice-Hall, Englevood
change, between 0.0 and 1.0, for the parameter is greater Cliffs, 1978.
than the requirement of 0.5, the parameter’s value change [19] H. Mehdizadeh, J.M. Dickson, Evaluation of Surface Force Pore
factor is computed. Flow and Modified Surface Force Pore Flow models for reverse
5. A multiplication change factor is randomly generated in osmosis transport, Chem. Eng. Comm. 103 (1991) 65–82.
a user-defined range, normally 0.7–1.3, for the parameter [20] H. Mehdizadeh, J.M. Dickson, P.K. Eriksson, Temperature effects on
the performance of thin-film composite aromatic polyamide mem-
under consideration. The new value of the parameter is branes, I & EC Res. 28 (1989) 814–824.
determined by multiplying the old one by this factor. [21] T. Matsuura, Y. Taketani, S. Sourirajan, Estimation of interfacial
6. The resulting values of simplex vertices and the set of forces governing the reverse-osmosis system: nonionized polar or-
parameters is accepted only if the error after the above ganic solute-water-cellulose acetate membrane, in: A.F. Turbak (Ed.),
described change is less than what it was before it car- Synthetic Membranes, ACS Symposium Series, American Chemical
Society, Washington, DC, 1981 (Chapter 19).
rying out the algorithm. [22] K. Chan, T. Matsuura, S. Sourirajan, Interfacial forces, average pore
size and pore size distribution of ultrafiltration membranes, Ind. Eng.
Chem. Prod. Res. Dev. 21 (4) (1982) 605–612.
References [23] S. Jain, Analysis of Modified Surface Force Pore Flow model with
concentration polarization and comparison with Spiegler–Kedem
model in reverse osmosis systems, M. Tech. thesis, I.I.T. Delhi, May
[1] L. Onsager, Reciprocal relations in irreversible processes. Part I,
2003.
Phys. Rev. 37 (1931) 405.
[24] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, Wi-
[2] L. Onsager, Reciprocal relations in irreversible processes. Part II,
ley, New York, 1960.
Phys. Rev. 38 (1931) 2265.
[25] S.T. Hwang, K. Kammermeyer, Membranes in Separation, Wiley,
[3] O. Kedem, A. Katchalsky, Thermodynamic analysis of the perme-
New York, 1975.
ability of biological membranes to non-electrolytes, Biochem. Bio-
[26] M.C. Porter, Concentration polarization with membrane ultrafilter-
phys. Acta 27 (1958) 229–246.
ation, Ind. Eng. Chem. Prod. Res. Dev. 11 (3) (1972) 234–248.
[4] O. Kedem, A. Katchalsky, Permeability of composite membranes.
[27] H. Mehdizadeh, J.M. Dickson, Solving non-linear differential equa-
Part 3. Series array of elements, Trans. Faraday Soc. 59 (1963)
tions of membrane transport by orthogonal collocation, Comput.
1941–1953.
Chem. Eng. 14 (2) (1990) 157–160.
[5] A. Katchalsky, P.F. Curran, Non-Equilibrium Thermodynamics in
[28] E.L. Cussler, Diffusion: Mass Transfer in Fluid Systems, Cambridge
Biophysics, Harvard University Press, Cambridge, MA, 1975.
University Press, Cambridge, 1984.
[6] K.S. Spiegler, O. Kedem, Thermodynamics of hyperfiltration (reverse
[29] A.H.P. Skelland, Diffusional Mass Transfer, Wiley/Interscience, USA,
osmosis): criteria for efficient membranes, Desalination 1 (1966) 311.
1974.
[7] H.K. Lonsdale, U. Merten, R.L. Riley, Transport properties of cel-
[30] J.A. Dean (Ed.), Lange’s Handbook of Chemistry, 15/e, McGraw
lulose acetate osmotic membranes, J. Appl. Polym. Sci. 9 (1965)
Hill, 1999, pp. 8.157–8.160.
1341–1362.
[31] W.L. McCabe, J.C. Smith, P. Harriott, Unit Operations of Chemical
[8] U. Merten (Ed.), Desalination by Reverse Osmosis, MIT Press,
Engineering, McGraw-Hill, 5th(Intl.)/e, 1993, pp. 1094–1102.
Cambridge, MA, 1966 (Chapter 2).
[32] Z.V.P. Murthy, S.K. Gupta, Estimation of mass transfer coefficient
[9] T.K. Sherwood, P.L.T. Brian, R.E. Fisher, Desalination by reverse
using a combined nonlinear membrane transport and film theory
osmosis, Ind. Eng. Chem. Fund. 6 (1) (1967) 2–12.
model, Desalination 109 (1997) 39–49.
[10] G. Jonsson, C.E. Boesen, Water and solute transport through cellulose
[33] J.A. Nelder, R. Mead, Simplex method for function minimization,
acetate reverse osmosis membranes, Desalination 17 (2) (1975) 145–
Comput. J. 7 (1965) 308–313.
165.
[34] J.L. Dellis, J.L. Carpentier, Nelder and Mead algorithm in impedance
[11] M. Soltanieh, W.N. Gill, Review of reverse osmosis membranes and
spectra fitting, Solid State Ionics 62 (1993) 119–123.
transport models, Chem. Eng. Comm. 12 (1981) 279.
[35] J.V. Beck, K.J. Arnold, Parameter Estimation in Engineering and
[12] E.A. Mason, H.K. Lonsdale, Statistical mechanical theory of mem-
Science, Wiley, NY, 1977.
brane transport, J. Membr. Sci. 51 (1990) 1.

You might also like