Está en la página 1de 395

Coupled modelling of hydrodynamics and mass transfer in

membrane filtration

by

Gregory Peter Keir


BE / BSc (Hons I)

Submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

Deakin University

August, 2012
DEAKIN UNIVERSITY
ACCESS TO THESIS  A

I am the author of the thesis entitled Coupled modelling of hydrodynamics and mass transfer in

membrane filtration submitted for the degree of Doctor of Philosophy.

This thesis may be made available for consultation, loan and limited copying in accordance with the
Copyright Act 1968.

'I certify that I am the student named below and that the information provided in the form is correct'

Full Name Gregory Peter Keir

Signed

Date 04/08/2012
DEAKIN UNIVERSITY
CANDIDATE DECLARATION

I certify that the thesis entitled Coupled modelling of hydrodynamics and mass
transfer in membrane filtration submitted for the degree of Doctor of
Philosophy is the result of my own work and that where reference is made to the
work of others, due acknowledgment is given.

I also certify that any material in the thesis which has been accepted for a degree

or diploma by any university or institution is identified in the text.

'I certify that I am the student named below and that the information provided in the form is correct'

Full Name Gregory Peter Keir

Signed

Date 04/08/2012
Abstract

The high-pressure membrane filtration processes of nanofiltration (NF) and

reverse osmosis (RO) are widely used in industry to remove dissolved

components from liquid streams, especially in the fields of water and wastewater

treatment. Despite their ubiquity, it is often difficult to predict the performance of

a membrane filtration system without expensive pilot studies. This is because the

flow of fluid through the membrane and the removal of dissolved components

depend both on the hydrodynamic conditions within the bulk of the membrane

module and the interactions between the dissolved components and the

microstructure of the membrane, which can be complex. This can be especially

apparent when multiple dissolved components are in solution.

This thesis outlines the development of a computational model to predict the

behaviour of such systems by combining a rigorous hydrodynamic model using

commercially available computational fluid dynamics (CFD) software with a

microscopic mass transfer model including the effects of convection, diffusion,

and electric migration to predict the removal of dissolved components by the

membrane. These fields of modelling in membrane applications have traditionally

been disparate; combining the two allows for a rigorous and fairly widely

applicable general predictive model of the NF and RO processes.

The CFD package ANSYS CFX is used for all simulations, allowing construction

of hydrodynamic models for arbitrary two-dimensional (2D) or three-dimensional

i
(3D) membrane channel geometries, and laminar or turbulent flow conditions.

The hydrodynamic model used here allows for the properties of the solution, such

as density, viscosity, solute diffusivity, and osmotic pressure to vary spatially

within the membrane channel depending on the local solute concentrations. This

allows for the formation of a concentration polarisation boundary layer near the

membrane surface, a commonly observed phenomenon which reduces the

effectiveness of NF and RO processes. In addition, the formulation presented here

allows for the permeate flux (the flow of fluid through the membrane) to vary

spatially depending on the local hydrodynamic conditions.

A substantial amount of additional program code has been written to couple the

CFD model to the microscopic mass transfer model, which is used to predict the

rejection of solutes and hence apply appropriate boundary conditions to the CFD

model. The mass transfer model is a refined version of the classic Donnan Steric

Partitioning Model (DSPM), which has been generalised to take molecular shape

effects into account for both charged and uncharged solutes. In addition, a method

has been described and implemented for numerical solution using the model for

an arbitrary combination of multiple charged and uncharged solutes. The

implementation of the coupled CFD-mass transfer model also allows for the

rejection of solutes to vary spatially throughout the membrane channel.

In summary, the novel research elements of this thesis are:

x Coupling of a commercial CFD code to a numerical solution of a refined

version of the DSPM to predict both membrane channel hydrodynamics

and mass transfer through NF and RO membranes;

ii
x Incorporation of both spatially variable permeate flux and solute rejection

into a CFD model of a membrane channel;

x Refinement of the classic DSPM to include molecular shape effects for

both charged and uncharged solutes; and

x Presentation of a numerical method of solution for the refined DSPM for

an arbitrary combination of multiple charged and uncharged solutes.

The coupled CFD-mass transfer model has been tested against experimental and

computational results from various sources, and found to give reliable results. In

addition, a series of numerical experiments have been carried out which highlight

some interesting phenomena predicted by the coupled model, including significant

spatial variation in predicted solute rejection, presence of an inverse concentration

polarisation layer for negatively rejected solutes, variation in near-wall

concentration gradients by solute, and significant variability in overall permeate

flux depending on solution composition. Finally, comments are made regarding

the suitability of the model to general application, and suggestions are made for

future research in extending the applicability of the model as computational

resources become more abundant.

iii
Acknowledgements

I would like to acknowledge the following for their support, hard work, and

guidance:

x Prof. Bas Baskaran, Mr. Mark Mitchell, Dr. Richard Yang, Mr. Steve

Bagshaw and others past and present within the School of Engineering at

Deakin University;

x Deakin University for providing a Deakin University Postgraduate

Research Scholarship;

x A/Prof. Siva Sivakugan, Prof. John Patterson, and others past and present

in the School of Engineering at James Cook University;

x The Graduate Research School at James Cook University for providing a

James Cook University Postgraduate Research Scholarship; and

x Dimuth Navaratna and my other fellow postgraduate students both at

Deakin University and James Cook University.

I must extend a special thank you to my thesis supervisor A/Prof. Jega

Jegatheesan for his tireless work, support, and guidance both as a supervisor and a

friend. I would also like to thank Dr. Li Shu for her guidance and patience.

Finally, and most importantly, I would like to thank my wife Merrin, and my

family, for their unconditional love and support.

iv
Table of contents

Abstract .................................................................................................................... i

Acknowledgements ................................................................................................ iv

Table of contents ..................................................................................................... v

List of figures .......................................................................................................... x

List of tables ....................................................................................................... xxvi

List of publications and presentations ................................................................ xxix

Chapter 1 Introduction............................................................................................. 1

1.1 Thesis outline ................................................................................................. 2

Chapter 2 Membrane filtration and the problem of flux decline ............................. 5

2.1 Membrane processes...................................................................................... 5

2.1.1 Membrane performance ........................................................................ 11

2.2 Membrane flux decline ................................................................................ 12

2.2.1 Concentration polarisation .................................................................... 13

2.2.2 Membrane fouling ................................................................................. 16

2.3 Methods to reduce flux decline.................................................................... 25

2.3.1 Pretreatment of feed solution ................................................................ 25

2.3.2 Modification of membrane properties................................................... 26

2.3.3 Control of membrane operating parameters.......................................... 27

2.3.4 Membrane cleaning ............................................................................... 35

2.4 Summary and directions for research .......................................................... 36

v
Chapter 3 The use of computational fluid dynamics in membrane filtration........ 38

3.1 Basic principles of CFD .............................................................................. 39

3.2 Modelling of membrane processes under laminar conditions ..................... 41

3.2.1 Flow through porous membranes.......................................................... 41

3.2.2 Concentration-polarisation effects ........................................................ 45

3.2.3 Cake formation effects .......................................................................... 50

3.2.4 The influence of osmotic pressure ........................................................ 55

3.3 Modelling of membrane processes under turbulent conditions ................... 58

3.3.1 Studies using Reynolds averaged Navier-Stokes methods ................... 59

3.3.2 Studies using direct numerical simulation methods.............................. 63

3.3.3 Summary of turbulent membrane models ............................................. 66

3.4 Prediction of membrane mass transfer ........................................................ 66

3.5 Comments on commercial membrane simulation packages........................ 74

3.6 Summary and directions for research .......................................................... 75

Chapter 4 Hydrodynamic modelling of membrane filtration ................................ 79

4.1 Introduction ................................................................................................. 79

4.2 2D flow through a porous membrane channel............................................. 80

4.2.1 Model geometry .................................................................................... 81

4.2.2 Boundary and initial conditions ............................................................ 82

4.2.3 Computational mesh, convergence and hydrodynamic verification ..... 84

4.2.4 Incorporation of concentration polarisation effects .............................. 86

4.2.5 Simulations ........................................................................................... 90

4.2.6 Concluding remarks ............................................................................ 103

4.3 Flow in spacer-filled channels ................................................................... 103

vi
4.3.1 Unsteady flow and turbulence in spacer-filled channels .................... 104

4.3.2 2D and 3D modelling .......................................................................... 105

4.3.3 Model geometry .................................................................................. 105

4.3.4 Boundary and initial conditions .......................................................... 107

4.3.5 Incorporation of concentration polarisation effects ............................ 108

4.3.6 Computational mesh and convergence criteria ................................... 109

4.3.7 Transient simulations .......................................................................... 111

4.3.8 Concluding remarks ............................................................................ 127

4.4 Additional remarks on 3D modelling ........................................................ 127

4.5 Conclusions................................................................................................ 134

Chapter 5 Prediction of membrane rejection ....................................................... 137

5.1 Introduction................................................................................................ 137

5.2 Approaches to describe separation in NF and RO processes .................... 138

5.3 Model theory and solution details.............................................................. 140

5.3.1 The extended Nernst-Planck equation ................................................ 140

5.3.2 Rejection of uncharged solutes ........................................................... 141

5.3.3 Rejection of charged solutes ............................................................... 147

5.3.4 Rejection of mixtures of charged and uncharged solutes ................... 151

5.3.5 A note on concentration polarisation and volumetric flux .................. 154

5.4 Computational and numerical considerations ............................................ 155

5.4.1 Use of existing numerical routines and libraries................................. 155

5.4.2 Program structure ................................................................................ 156

5.4.3 Determination of Donnan partitioning equilibrium constants ............ 157

5.5 Results and discussion ............................................................................... 158

vii
5.5.1 Data from Bowen and Mohammad ..................................................... 159

5.5.2 Data from Bowen and Mukhtar .......................................................... 167

5.5.3 Predictions for solutions with four or more charged solutes .............. 178

5.6 Conclusions ............................................................................................... 182

Chapter 6 A fully predictive membrane filtration model: coupled hydrodynamic


and predictive rejection modelling ...................................................................... 184

6.1 Introduction ............................................................................................... 184

6.2 Coupled modelling of solutions containing uncharged solutes ................. 185

6.2.1 Model details ....................................................................................... 185

6.2.2 Model verification ............................................................................... 188

6.2.3 Simulations for low transmembrane pressure ..................................... 193

6.3 Coupled modelling of solutions containing charged solutes ..................... 197

6.3.1 Model details ....................................................................................... 198

6.3.2 Special considerations for the coupled model .................................... 202

6.3.3 Model verification – two charged solutes ........................................... 203

6.3.4 Simulations – multiple charged solutes .............................................. 212

6.4 Coupled modelling of mixed solutions...................................................... 235

6.4.1 Model details ....................................................................................... 235

6.4.2 Simulations ......................................................................................... 238

6.5 Conclusions ............................................................................................... 246

Chapter 7 Simulations using the coupled hydrodynamic and rejection model ... 251

7.1 Introduction ............................................................................................... 251

7.2 Simulation approach .................................................................................. 252

7.2.1 Geometry............................................................................................. 253

viii
7.2.2 Boundary and initial conditions .......................................................... 254

7.2.3 Computational mesh and convergence criteria ................................... 255

7.3 Coupled simulation of single salt solution in spacer filled channel .......... 256

7.4 Coupled simulation of multiple salt solution in spacer filled channel....... 267

7.5 Coupled simulation of complex mixed solution in spacer filled channel .. 286

7.6 Conclusions................................................................................................ 300

Chapter 8 Conclusions......................................................................................... 304

8.1 Overview.................................................................................................... 304

8.2 Hydrodynamic modelling – conclusions and recommendations ............... 305

8.3 Mass transfer modelling for rejection prediction – conclusions and


recommendations ............................................................................................. 308

8.4 Coupled hydrodynamic-mass transfer modelling – conclusions and


recommendations ............................................................................................. 311

Appendix A Sample MATLAB numerical codes for predictive rejection model


............................................................................................................................. 316

Appendix B Sample FORTRAN numerical codes for coupled rejection model –


User Routine approach ........................................................................................ 334

ix
List of figures

Figure 2.1 - Schematic representation of: (a) typical membrane filtration process,

and (b) typical two-phase separation via membrane [after 2]. ................................ 7

Figure 2.2 - Typical flux decline during filtration run [after 2] ............................ 13

Figure 2.3 - Concentration polarisation profile under steady-state conditions [after

2]............................................................................................................................ 14

Figure 2.4 - Typical net-type spacer geometry, showing (a) use in spiral-wound

membrane [47], and (b) woven and non-woven spacers [35]. .............................. 30

Figure 2.5 - Flow regimes for gas-injected flow through membrane [35] ............ 31

Figure 2.6 - Helical screw thread inserts used to induce Dean vortices (left: one

start thread, right: three start thread) [57] ............................................................. 32

Figure 2.7 - Particle detachment mechanisms for ultrasonic agitation [68].......... 35

Figure 3.1 - Finite element mesh used to model concentration polarisation

phenomena [90] ..................................................................................................... 47

Figure 3.2 - Solution algorithm proposed by Richardson and Nassehi [91] ......... 49

Figure 3.3 - Solution algorithm for Lee and Clark [93] cake formation model .... 54

Figure 3.4 - Deviation of analytical and numerical velocity profiles for Pellerin et

al.'s model for (a) ultrafiltration, (b) microfiltration [5]. ....................................... 61

Figure 3.5 – Difference in ionic concentrations predicted in concentration

polarisation boundary layer using both Nernst-Planck and Fickian diffusion

models [103] .......................................................................................................... 73

x
Figure 4.1 - Model geometry for basic 2D model of laminar flow through porous

membrane channel (not to scale) ........................................................................... 82

Figure 4.2 - Comparison of dimensionless axial pressure drop for CFD model

with analytical solutions for hydrodynamic verification of model ....................... 86

Figure 4.3 - Representative detail of computational mesh for 2D channel with

refinement for concentration polarisation boundary layer at (a) channel inlet, (b)

membrane surface (not to scale) ............................................................................ 89

Figure 4.4 - Comparison of predicted permeate fluxes for CFD model with

experimental and modelled fluxes from [94] using experimentally determined

rejection factors ..................................................................................................... 93

Figure 4.5 - Comparison of dimensionless CP boundary layer profile for CFD

model with computational results of [94] using experimentally determined

rejection factors ..................................................................................................... 94

Figure 4.6 - Representative detail of variation in solute mass fraction in 2D

membrane channel including concentration polarisation boundary layer (sucrose

solution, TMP 1 MPa, NRe = 500 – not to scale) ................................................... 96

Figure 4.7 - Representative detail of velocity variation at inlet of 2D membrane

channel including concentration polarisation boundary layer (sucrose solution,

TMP 1 MPa, NRe = 500 – not to scale) .................................................................. 96

Figure 4.8 - Representative detail of velocity variation for entire 2D membrane

channel including concentration polarisation boundary layer (sucrose solution,

TMP 1 MPa, NRe = 500 – not to scale) .................................................................. 97

xi
Figure 4.9 - Representative detail of density variation for entire 2D membrane

channel including concentration polarisation boundary layer (sucrose solution,

TMP 1 MPa, NRe = 500 – not to scale) .................................................................. 97

Figure 4.10 - Representative detail of viscosity variation for entire 2D membrane

channel including concentration polarisation boundary layer (sucrose solution,

TMP 1 MPa, NRe = 500 – not to scale) .................................................................. 98

Figure 4.11 - Representative detail of osmotic pressure variation for entire 2D

membrane channel including concentration polarisation boundary layer (sucrose

solution, TMP 1 MPa, NRe = 500 – not to scale) ................................................... 98

Figure 4.12 - Comparison of predicted permeate fluxes for CFD model

incorporating osmotic pressure correction factor with experimental and modelled

fluxes from [94] using experimentally determined rejection factors .................. 101

Figure 4.13 - Comparison of dimensionless CP boundary layer profile for CFD

model incorporating osmotic pressure correction factor with computational results

of [94] using experimentally determined rejection factors ................................. 102

Figure 4.14 – Model geometry for simulations of flow through 2D space-filled

channel ................................................................................................................ 106

Figure 4.15 – Representative detail of computational mesh for 2D spacer-filled

channel CFD model ............................................................................................. 110

Figure 4.16 – Comparison of predicted axial pressure drop as a function of

channel Reynolds number against computational results of [115] for spacer-filled

channel ................................................................................................................ 113

xii
Figure 4.17 – Comparison of predicted area-averaged wall shear stress as a

function of axial pressure drop against computational results of [115] for spacer-

filled channel ....................................................................................................... 114

Figure 4.18 – Comparison of predicted area-averaged salt mass fraction as a

function of channel Reynolds number against computational results of [115] for

spacer-filled channel............................................................................................ 114

Figure 4.19 – Predicted salt mass fraction distribution for spacer-filled channel for

conditions in [115] for steady flow at lower channel Reynolds numbers ........... 116

Figure 4.20 – Predicted fluid streamlines for spacer-filled channel for conditions

in [115] for steady flow at lower channel Reynolds numbers ............................. 117

Figure 4.21 – Predicted velocity vectors for spacer-filled channel for conditions in

[115] for steady flow at lower channel Reynolds numbers ................................. 118

Figure 4.22 – Predicted vorticity fields for spacer-filled channel for conditions in

[115] for steady flow at lower channel Reynolds numbers ................................. 119

Figure 4.23 – Predicted salt mass fraction distribution for spacer-filled channel for

conditions in [115] for unsteady flow at high channel Reynolds number (Rech =

800) at successive 6.1 ms timesteps (top to bottom) ........................................... 123

Figure 4.24 – Predicted fluid streamlines for spacer-filled channel for conditions

in [115] for unsteady flow at high channel Reynolds number (Rech = 800) at

successive 6.1 ms timesteps (top to bottom) ....................................................... 124

Figure 4.25 – Predicted velocity vectors for spacer-filled channel for conditions in

[115] for unsteady flow at high channel Reynolds number (Rech = 800) at

successive 6.1 ms timesteps (top to bottom) ....................................................... 125

xiii
Figure 4.26 – Predicted vorticity fields for spacer-filled channel for conditions in

[115] for unsteady flow at high channel Reynolds number (Rech = 800) at

successive 6.1 ms timesteps (top to bottom) ....................................................... 126

Figure 4.27 – Quarter unit-cell geometry corresponding to 2D model dimensions

used in [115] ........................................................................................................ 130

Figure 4.28 – Coarse 3D mesh (9.7 million elements) for unit-cell corresponding

to 2D model dimensions used in [115]................................................................ 131

Figure 4.29 – Representative detail of coarse 3D mesh (9.7 million elements) for

unit-cell corresponding to 2D model dimensions used in [115] ......................... 131

Figure 5.1 – Flowchart of calculation procedure for combination of charged and

uncharged solutes ................................................................................................ 153

Figure 5.2 - Comparison of experimental rejection data [141] and numerically

computed rejections as a function of flux ........................................................... 161

Figure 5.3 - Comparison of experimental rejection data [141] and numerically

computed rejections as a function of flux for single salt solutions at fixed

concentration ....................................................................................................... 166

Figure 5.4 - Comparison of experimental rejection data [141] and numerically

computed rejections as a function of flux for sodium chloride at varying

concentrations...................................................................................................... 166

Figure 5.5 – Variation of charge density as a function of bulk concentration using

experimental data from [141] .............................................................................. 167

Figure 5.6 - Comparison of experimental rejection data from Bowen and Mukhtar

[17] and numerically computed rejections for PES5 membrane for 10-3 M sodium

FKORULGHVROXWLRQXVLQJǻy/Ak and CX as fitting parameters ................................ 170

xiv
Figure 5.7 - Comparison of experimental rejection data from Bowen and Mukhtar

[17] and numerically computed rejections for PES5 membrane for 10-3 M sodium

FKORULGHVROXWLRQXVLQJǻy/Ak, CX, and rp as fitting parameters........................... 171

Figure 5.8 - Comparison of experimental rejection data from Bowen and Mukhtar

[17] and numerically computed rejections for PES5 membrane for 3×10-3 M

sodium sulphate solution using CX as fitting parameter ...................................... 173

Figure 5.9 - Variation of charge density as a function of bulk concentration using

experimental data from [17] ................................................................................ 175

Figure 5.10 - Comparison of experimental rejection data from Bowen and

Mukhtar [17] and numerically computed rejections for PES5 membrane for 1×10-3

M salt solution using membrane charge density isotherm for non-spherical pore

model with molar ratios of: (a) 0.2, (b) 0.4, (c) 0.6, (d) 0.8 (NaCl:Na2SO4) ...... 177

Figure 5.11 - Numerically computed rejections for PES5 membrane for 1×10-3 M

salt solution using membrane charge density isotherm for non-spherical pore

model with molar ratios of: (a) 0.333:0.333:0.333, (b) 0.25:0.25:0.5, (c)

0.25:0.5:0.25, (d) 0.5:0.25:0.25 (NaCl:Na2SO4:CaCl2)....................................... 179

Figure 5.12 - Numerically computed rejections for PES5 membrane for 1×10-3 M

salt solution using membrane charge density isotherm for non-spherical pore

model with molar ratios as listed in Table 5.11 .................................................. 181

Figure 6.1 – Flowchart of computation procedure for coupled hydrodynamic-mass

transfer model for uncharged solutes only .......................................................... 187

Figure 6.2 - Comparison of predicted permeate fluxes for CFD model with

experimentally determined fluxes of [94] using calculated spatially varying

rejection factors from pore model ....................................................................... 192

xv
Figure 6.3 - Comparison of dimensionless CP boundary layer profile for CFD

model with computational results of [94] using calculated spatially varying

rejection factors from pore model ....................................................................... 193

Figure 6.4 - Axial variation in sucrose rejection for low transmembrane pressure

with spatially varying rejection factor ................................................................. 194

Figure 6.5 - Comparison of predicted permeate fluxes for CFD model for low

transmembrane pressures using calculated constant and spatially varying rejection

factors .................................................................................................................. 196

Figure 6.6 - Comparison of dimensionless CP boundary layer profile for CFD

model for low transmembrane pressures using calculated constant and spatially

varying rejection factors ...................................................................................... 196

Figure 6.7 – Flowchart of computation procedure for coupled hydrodynamic-mass

transfer model for charged solutes only .............................................................. 201

Figure 6.8 – Comparison of predicted rejection and experimental rejection for

NaCl solution with CDNF50l membrane using fitted parameters from Table 6.5

............................................................................................................................. 207

Figure 6.9 – Comparison of experimental rejection (unfilled markers), predicted

rejection from pore model (solid line) and predicted rejection from coupled CFD-

pore model (filled markers) for NaCl solution with CDNF50l membrane using

fitted parameters from Table 6.5 ......................................................................... 209

Figure 6.10 – Predicted axial rejection variation from coupled CFD-pore model

for NaCl solution with CDNF50l membrane using fitted parameters from Table

6.5 ........................................................................................................................ 210

xvi
Figure 6.11 - Comparison of dimensionless CP boundary layer profile for coupled

CFD-pore model with computational results of [94] .......................................... 211

Figure 6.12 - Comparison of experimental rejection data from Bowen and

Mukhtar [17], numerically computed rejections from pore model, and area-

averaged predicted rejections from coupled hydrodynamic / pore model for PES5

membrane for 1×10-3 M salt solution with molar ratios of: (a) 0.2, (b) 0.4, (c) 0.6,

(d) 0.8 (NaCl:Na2SO4) ......................................................................................... 216

Figure 6.13 – Predicted axial rejection variation for PES5 membrane for 1×10-3 M

salt solution with molar ratios of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa ........ 218

Figure 6.14 – Predicted axial permeate flux variation for PES5 membrane for

1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4) ........................... 219

Figure 6.15 – Flowchart of computation procedure for coupled hydrodynamic-

mass transfer model for multiple charged solutes with simplified diffusion model

............................................................................................................................. 224

Figure 6.16 - Comparison of experimental rejection data from Bowen and

Mukhtar [17], numerically computed rejections from pore model, and area-

averaged predicted rejections from coupled hydrodynamic / pore model both with

and without concentration polarisation effects for PES5 membrane for 1×10-3 M

salt solution with molar ratio of 0.2 (NaCl:Na2SO4) ........................................... 228

Figure 6.17 – Predicted axial rejection variation for PES5 membrane both with

and without concentration polarisation effects for 1×10-3 M salt solution with

molar ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa ...................................... 230

xvii
Figure 6.18 – Predicted axial permeate flux variation for PES5 membrane both

with and without concentration polarisation effects for 1×10-3 M salt solution with

molar ratio of 0.2 (NaCl:Na2SO4) ....................................................................... 231

Figure 6.19 – Detail of computed concentration fields (normalised relative to bulk

inlet concentration) of (a) Na+, (b) Cl-, (c) SO42- for PES5 membrane with

concentration polarisation effects for 1×10-3 M salt solution with molar ratio of

0.2 (NaCl:Na2SO4) for TMP of 200 kPa ............................................................. 233

Figure 6.20 – Predicted variation normal to membrane surface for solute

concentration normalised relative to inlet bulk concentration for PES5 membrane

with concentration polarisation effects for 1×10-3 M salt solution with molar ratio

of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa ......................................................... 234

Figure 6.21 – Flowchart of computation procedure for coupled hydrodynamic-

mass transfer model for multiple charged and/or uncharged solutes with simplified

diffusion model ................................................................................................... 237

Figure 6.22 - Comparison of area-averaged predicted rejections from coupled

hydrodynamic / pore model with concentration polarisation effects for PES5

membrane for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4) with

and without sucrose addition ............................................................................... 240

Figure 6.23 – Predicted axial rejection variation for PES5 membrane for 1×10-3 M

salt solution with molar ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa with and

without sucrose addition...................................................................................... 243

Figure 6.24 – Predicted axial permeate flux variation for PES5 membrane both

with and without sucrose addition for 1×10-3 M salt solution with molar ratio of

0.2 (NaCl:Na2SO4) .............................................................................................. 244

xviii
Figure 6.25 – Detail of computed concentration fields (normalised relative to inlet

bulk concentration) of (a) Na+, (b) Cl-, (c) SO42-, (d) sucrose for PES5 membrane

for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4) with sucrose

addition for TMP of 200 kPa............................................................................... 245

Figure 6.26 – Predicted variation normal to membrane surface for solute

concentration normalised relative to inlet bulk concentration for PES5 membrane

for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4) with sucrose

addition for TMP of 200 kPa............................................................................... 246

Figure 7.1 – Model geometry for simulations of flow through 2D space-filled

channel................................................................................................................. 254

Figure 7.2 – Predicted salt mass fraction distribution for spacer-filled channel

using coupled-mass transfer model for single salt solution (NaCl, 0.002 kg kg-1,

equivalent to 34.1 mol m-3) ................................................................................. 257

Figure 7.3 – Predicted fluid streamlines for spacer-filled channel using coupled-

mass transfer model for single salt solution (NaCl, 0.002 kg kg-1, equivalent to

34.1 mol m-3) ....................................................................................................... 258

Figure 7.4 – Predicted velocity vectors for spacer-filled channel using coupled-

mass transfer model for single salt solution (NaCl, 0.002 kg kg-1, equivalent to

34.1 mol m-3) ....................................................................................................... 258

Figure 7.5 – Predicted vorticity fields for spacer-filled channel using coupled-

mass transfer model for single salt solution (NaCl, 0.002 kg kg-1, equivalent to

34.1 mol m-3) ....................................................................................................... 259

xix
Figure 7.6 - Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection (NaCl, 0.002 kg

kg-1, equivalent to 34.1 mol m-3, Rech = 100) ...................................................... 261

Figure 7.7 - Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection (NaCl, 0.002 kg

kg-1, equivalent to 34.1 mol m-3, Rech = 200) ...................................................... 262

Figure 7.8 - Predicted transverse variation in normalised solute concentration as a

function of non-dimensionalised distance from lower membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3) .............................................. 265

Figure 7.9 - Predicted transverse variation in normalised solute concentration as a

function of non-dimensionalised distance from upper membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3) .............................................. 265

Figure 7.10 – Predicted ionic mass fraction distribution for spacer-filled channel

using coupled-mass transfer model for multiple salt solution (NaCl:Na2SO4 with

molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3), Rech = 100 ................... 268

Figure 7.11 – Predicted ionic mass fraction distribution for spacer-filled channel

using coupled-mass transfer model for multiple salt solution (NaCl:Na2SO4 with

molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3), Rech = 200 ................... 269

Figure 7.12 – Predicted fluid streamlines for spacer-filled channel using coupled-

mass transfer model for multiple salt solution (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3) ......................................................... 270

xx
Figure 7.13 – Predicted velocity vectors for spacer-filled channel using coupled-

mass transfer model for multiple salt solution (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3) ......................................................... 270

Figure 7.14 – Predicted vorticity fields for spacer-filled channel using coupled-

mass transfer model for multiple salt solution (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3) ......................................................... 271

Figure 7.15 - Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech = 100) ........... 275

Figure 7.16 - Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech = 200) ........... 276

Figure 7.17 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from lower membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech =

100)...................................................................................................................... 278

Figure 7.18 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from lower membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech =

200)...................................................................................................................... 278

Figure 7.19 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from upper membrane wall within

xxi
spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech =

100)...................................................................................................................... 279

Figure 7.20 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from upper membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech =

200)...................................................................................................................... 279

Figure 7.21 – Difference in predicted ionic mass fraction distribution for spacer-

filled channel between constant rejection model and variable rejection model for

multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic

strength 34.1 eq m-3), Rech = 100 ........................................................................ 281

Figure 7.22 – Difference in predicted ionic mass fraction distribution for spacer-

filled channel between constant rejection model and variable rejection model for

multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic

strength 34.1 eq m-3), Rech = 200 ........................................................................ 283

Figure 7.23 – Difference in predicted permeate flux within spacer-filled channel w

between constant rejection model and variable rejection model (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech = 100) ........... 284

Figure 7.24 – Difference in predicted permeate flux within spacer-filled channel w

between constant rejection model and variable rejection model (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, Rech = 200) ........... 285

Figure 7.25 – Predicted ionic mass fraction distribution for spacer-filled channel

using coupled-mass transfer model for complex mixture solution (NaCl:Na2SO4

xxii
with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose

at 0.001 kg kg-1), Rech = 100 ............................................................................... 287

Figure 7.26 – Predicted ionic mass fraction distribution for spacer-filled channel

using coupled-mass transfer model for complex mixture solution (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose

at 0.001 kg kg-1), Rech = 200 ............................................................................... 288

Figure 7.27 – Predicted fluid streamlines for spacer-filled channel using coupled-

mass transfer model for complex mixture solution (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1)

............................................................................................................................. 289

Figure 7.28 – Predicted velocity vectors for spacer-filled channel using coupled-

mass transfer model for complex mixture solution (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1)

............................................................................................................................. 289

Figure 7.29 – Predicted vorticity fields for spacer-filled channel using coupled-

mass transfer model for single complex mixture solution (NaCl:Na2SO4 with

molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at

0.001 kg kg-1) ...................................................................................................... 290

Figure 7.30 – Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4:sucrose solution, Rech = 100) ...................................................... 292

Figure 7.31 - Predicted variation in mass transfer parameters within spacer-filled

channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4:sucrose solution, Rech = 200) ...................................................... 293

xxiii
Figure 7.32 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from lower membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1, Rech = 100) ................................................. 294

Figure 7.33 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from lower membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1, Rech = 200) ................................................. 294

Figure 7.34 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from upper membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1, Rech = 100) ................................................. 295

Figure 7.35 - Predicted transverse variation in normalised solute concentration as

a function of non-dimensionalised distance from upper membrane wall within

spacer-filled channel with spatially varying permeate flux and solute rejection

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1, Rech = 200) ................................................. 295

Figure 7.36 – Difference in predicted solute mass fraction distribution for spacer-

filled channel between constant rejection model and variable rejection model for

complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic

strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1), Rech = 100 ........ 297

xxiv
Figure 7.37 – Difference in predicted solute mass fraction distribution for spacer-

filled channel between constant rejection model and variable rejection model for

complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic

strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1), Rech = 200......... 298

Figure 7.38 – Difference in predicted permeate flux within spacer-filled channel w

between constant rejection model and variable rejection model (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose

at 0.001 kg kg-1, Rech = 100) ............................................................................... 299

Figure 7.39 – Difference in predicted permeate flux within spacer-filled channel w

between constant rejection model and variable rejection model (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose

at 0.001 kg kg-1, Rech = 200) ............................................................................... 299

xxv
List of tables

Table 2.1 – Summary of common membrane processes [after 2] ........................... 8

Table 2.2 - Comparison of pressure-driven membrane processes [after 2] ............ 9

Table 2.3 - Summary of models derived for constant pressure filtration law for

fouled membranes [3]............................................................................................ 21

Table 2.4 - Summary of UF filtration models for fouled membranes [compiled

from 3] ................................................................................................................... 22

Table 2.5 - Summary of RO filtration models [compiled from 3] ........................ 24

Table 3.1- Summary of characteristics of discretisation methods [compiled from

79].......................................................................................................................... 40

Table 4.1 - Model input parameters for verification case ..................................... 85

Table 4.2 - Trial solute components and expressions for transport properties (after

[94]) ....................................................................................................................... 88

Table 4.3 – Model input parameters for simulations corresponding with Geraldes

et al. [94] ............................................................................................................... 90

Table 4.4 - Experimentally observed rejection factors for test compounds from

[94] for NRe = 2000 ................................................................................................ 91

Table 4.5 - Osmotic pressure correction factor expressions for test compounds

(after [94]) ............................................................................................................. 99

Table 4.6 – Model input parameters for simulations corresponding with Wardeh &

Morvan [115]....................................................................................................... 108

xxvi
Table 5.1 – Summary of existing numerical libraries and routines used in

development of predictive rejection model code ................................................ 155

Table 5.2 - Physical properties for uncharged solutes used in [141] .................. 160

Table 5.3 – Fitting results for uncharged solutes used in [141] .......................... 161

Table 5.4 - Physical properties for ionic solutes used in [141] ........................... 164

Table 5.5 – Fitting results for single salt solutions used in [141] ....................... 165

Table 5.6 - Physical properties for solutes from [17] .......................................... 169

Table 5.7 - Comparison of fitting for ǻy/Ak and CX for PES5 membrane used by

Bowen and Mukhtar [17] for 10-3 M sodium chloride solution .......................... 170

Table 5.8 - Comparison of ILWWLQJIRUǻy/Ak, CX, and rp for PES5 membrane used

by Bowen and Mukhtar [17] for 10-3 M sodium chloride solution ..................... 172

Table 5.9 - Comparison of fitting for CX for PES5 membrane used by Bowen and

Mukhtar [17] for 3×10-3 M sodium sulphate solution ......................................... 173

Table 5.10 – Fitted parameters for charge density isotherms using experimental

data from [17] ...................................................................................................... 175

Table 5.11 – Molar ratios used for simulation of six ion 1×10-3 M salt solution 180

Table 6.1 - Molecular properties for uncharged test compounds derived using

MOPAC2009 software ........................................................................................ 188

Table 6.2 - Predicted and experimentally observed rejection factors for test

compounds for CDNF50l thin-film composite NF membrane (Separem, Biela,

Italy) .................................................................................................................... 191

Table 6.3 – Minimum arguments to be passed from CFX solver to external

Fortran routine for calculating rejection of charged solutes................................ 198

xxvii
Table 6.4 – Estimates for permeate fluxes and wall concentrations for NaCl

solution based on previous CFD results .............................................................. 204

Table 6.5 – Fitted membrane parameters for CDNF50l membrane for NaCl

solution ................................................................................................................ 206

Table 6.6 – Model input parameters for simulations corresponding with Bowen &

Mukhtar [17] ....................................................................................................... 214

Table 6.7 – Properties for additional uncharged solute (sucrose) in coupled

simulation of charged-uncharged solute mixture ................................................ 238

Table 7.1 – Model input parameters for coupled simulations of 2D spacer-filled

channel ................................................................................................................ 255

Table 7.2 – Comparison of average permeate flux values for single salt (NaCl,

overall ionic strength 34.1 eq m-3) and multiple salt (NaCl:Na2SO4 with molar

ratio 0.2:0.8, overall ionic strength 34.1 eq m-3) cases........................................ 273

Table 7.3 – Fixed rejection values used for uncoupled model comparison case

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3) ...... 280

Table 7.4 – Comparison of average permeate flux values for single salt (NaCl,

overall ionic strength 34.1 eq m-3), multiple salt (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3), and complex mixture (NaCl:Na2SO4

with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose

at 0.001 kg kg-1) cases ......................................................................................... 290

Table 7.5 – Fixed rejection values used for uncoupled model comparison case

(NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1) .................................................................... 296

xxviii
List of publications and presentations

The following publications were derived from the work contained in this thesis:

G. Keir, V. Jegatheesan. Prediction of solute rejection and modelling of steady-

state concentration polarisation effects in pressure-driven membrane filtration

using computational fluid dynamics, Membrane Water Technology, Vol.3, No.2,

(2012).

The following presentations were derived from the work contained in this thesis:

G. Keir, V. Jegatheesan (2011), Prediction of solute rejection and modelling of

steady-state concentration polarisation effects in pressure-driven membrane

filtration using computational fluid dynamics, Conferences on the Challenges in

Environmental Science and Engineering 2011, Tainan, Taiwan.

G. Keir, V. Jegatheesan (2011), Modelling of nanofiltration and reverse osmosis

using coupled hydrodynamic and pore models, Institute for Technology Research

and Innovation Research Conference 2011, Geelong, Victoria.

xxix
G. Keir, V. Jegatheesan (2011), Modelling of Fouling Processes in Membrane

Filtration Using Computational Fluid Dynamics, Victorian Modelling Group

Meeting #6 for the Australian Hydraulic Modelling Association, Geelong,

Victoria.

xxx
Chapter 1

Introduction

Membrane filtration is a common industrial process for removing dissolved

components from liquid streams, especially in water and wastewater treatment.

While the process is widely used, predicting the behaviour and performance of a

particular membrane system is difficult. This is due to both the complex

hydrodynamic behaviour of the bulk flow which passes through membrane

modules, and the complex interactions between the dissolved constituents and the

membrane microstructure. Because of this, performance can vary widely based on

feed stream composition and operating conditions, and hence to determine

suitability for a particular application, pilot-scale studies are often required.

Development of a general predictive model for membrane filtration processes

would therefore be of significant benefit.

However, most research into membrane modelling has considered the

hydrodynamic processes within the membrane channel, and the mass transfer

processes governing removal of dissolved components by the membrane, in

isolation. The aim of this study is to reconcile these two approaches, and develop

an integrated computational model to describe both the membrane channel

hydrodynamics using a commercial computational fluid dynamics (CFD) code,

and the membrane mass transfer with appropriate numerical models.

1
1.1 Thesis outline

This thesis consists of eight chapters which are described in the following.

Chapter 1 (the present section) outlines the structure of the document, and

provides a concise statement of the research question and arguments for the

relevance and importance of the study.

Chapter 2 presents an overview of pressure-driven membrane filtration processes,

including microfiltration, ultrafiltration, nanofiltration, and reverse osmosis, and

describes the problem of flux decline inherent to all these processes. It describes

the mathematical models which have been developed to model these processes, as

well as presenting techniques that have been proposed to reduce or minimise this

flux decline.

Chapter 3 describes the basic principles of computational fluid dynamics, and

investigates the use of computational fluid dynamics techniques in describing the

membrane filtration process. The modelling of concentration polarisation effects

and membrane fouling is detailed in order to describe the process of membrane

flux decline. As well, the use of transient simulations and numerical turbulence

models is investigated.

Chapter 4 describes the development of a basic set of rigorous hydrodynamic

models of the membrane filtration process. These models are validated against

experimental and computational results presented in the literature, and provide a

2
useful method of predicting membrane hydrodynamic performance in the cases

where the rejection behaviour of the membrane is well known. A variety of

membrane channel geometric configurations are investigated, and relevant

hydrodynamic phenomena are discussed.

Chapter 5 presents a numerical method which has been developed in order to

predict the rejection of various components in solution by nanofiltration and

reverse osmosis membranes. The method, unlike most other methods presented in

the literature, is general enough to predict the rejection for all elements of a

complex mixture composed of arbitrary numbers of charged and uncharged

solutes. The method has been developed in order to be coupled with a rigorous

hydrodynamic simulation of the membrane filtration process as described in

Chapter 4, which allows for a complete predictive model of the membrane

filtration process to be established.

Chapter 6 describes the coupling of the hydrodynamic models presented in

Chapter 4 and the predictive rejection models of Chapter 5 to present a fully

predictive model of the membrane filtration process for nanofiltration and reverse

osmosis membranes. The model is verified against experimental and

computational results from the literature, and observations are made regarding the

implementation of the model in terms of numerical stability, general utility, and

areas for future refinement.

3
In Chapter 7, the coupled hydrodynamic-mass transfer model is used to perform

detailed simulations of realistic membrane channel geometries and hydrodynamic

conditions. A series of numerical experiments are made using the new model, and

salient hydrodynamic and mass transfer phenomena are identified and discussed.

Chapter 8 summarises the findings made in Chapter 4, Chapter 5, Chapter 6, and

Chapter 7, and presents a set of conclusions regarding the use of coupled

hydrodynamic and predictive rejection models for membrane filtration.

Comments are made regarding the suitability of this computational approach for

predicting the overall behaviour of membrane filtration systems in terms of

computational efficiency and general applicability. Finally, several suggestions

are made for future work in this area.

4
Chapter 2

Membrane filtration and the problem of flux decline

Membrane filtration plays an important role in modern life, and has grown in a

matter of years from a specialised laboratory tool into a widespread industrial

process with significant industrial, economic and environmental importance [1].

Membranes are used today for a variety of large scale applications, including:

x treatment of drinking water to remove contaminants and pathogens;

x production of potable water from seawater;

x treatment of domestic and industrial wastewater;

x recovery of valuable constituents from industrial effluent; and

x various medical applications, such as removal of urea and toxins from blood

streams by dialysis.

While there are a multitude of different membrane processes and applications,

there are certain underlying features common to all membrane processes, as

discussed in the following.

2.1 Membrane processes

Membrane processes typically consist of a feed stream consisting of several

phases (solvents and solute), which is then divided by a membrane module into a

5
permeate stream and a concentrate stream [2]. A membrane can be considered in

the general sense as a barrier which separates two phases and restricts the

transport of various chemical species in a specific manner [1]. Based on this

definition, a great number of materials and structures may be considered as

membranes. The primary method of classifying membranes is the distinction

between biological and synthetic membranes, though nearly all membranes of any

industrial importance are synthetic.

Mulder [2] identified three basic types of synthetic membranes:

x porous membranes – these separate phases by discriminating between particle

sizes;

x non-porous membranes – these separate phases through differences in

solubility and/or diffusivity, and can separate particles of similar size; and

x carrier membranes – these separate phases by use of very specific carrier

molecules which facilitate transport of specific components.

Separation characteristics of a membrane are largely determined by the

dimensions of the membrane structure for porous membranes, while selection for

other membrane types is mostly dictated by the material properties of the

membrane. However, it should be emphasised that there is no discrete transition

between membrane types, and some processes may be intermediate between these

groupings.

6
Transport through the membrane occurs as a result of some driving force, such as

a gradient in pressure, temperature, concentration or electric potential. A

schematic representation of the membrane process is shown in Figure 2.1. The

driving force for the membrane process varies depending on the specific

application. A list of common membrane processes and their driving mechanisms

is shown in Table 2.1.

membrane

feed stream concentrate


membrane
module

feed permeate
permeate driving
force

(a) (b)

Figure 2.1 - Schematic representation of: (a) typical membrane filtration process, and (b) typical two-phase

separation via membrane [after 2].

The membrane processes generally of interest to the environmental engineer, and

those that will be considered in this thesis, are the conventional pressure-driven

membrane processes. These are (in order of decreasing size of constituents

treated):

x microfiltration (MF);

x ultrafiltration (UF);

7
x nanofiltration (NF); and

x reverse osmosis (RO).

A comparison of typical parameters and applications for these processes is given

in Table 2.2.

Table 2.1 – Summary of common membrane processes [after 2]

Membrane process Phase 1 Phase 2 Driving force

Microfiltration Liquid Liquid Hydraulic pressure

Ultrafiltration Liquid Liquid Hydraulic pressure

Nanofiltration Liquid Liquid Hydraulic pressure

Reverse osmosis Liquid Liquid Hydraulic pressure

Piezodialysis Liquid Liquid Hydraulic pressure

Gas separation Gas Gas Vapour pressure

Vapour permeation Gas Gas Vapour pressure

Pervaporation Liquid Gas Vapour pressure

Electrodialysis Liquid Liquid Electrical potential

Membrane electrolysis Liquid Liquid Electrical potential

Dialysis Liquid Liquid Concentration

Diffusion dialysis Liquid Liquid Concentration

Membrane contactors Liquid Liquid Concentration

Gas Liquid Concentration / vapour pressure

Liquid Gas Concentration / vapour pressure

Thermo-osmosis Liquid Liquid Temperature / vapour pressure

Membrane distillation Liquid Liquid Temperature / vapour pressure

8
Table 2.2 - Comparison of pressure-driven membrane processes [after 2]

Characteristic Microfiltration Ultrafiltration Nanofiltration Reverse osmosis

Membrane type Symmetric/asymmetric Asymmetric Composite Asymmetric,


porous porous composite

Thickness 10-150 μm 150 μm Sublayer 150 μm, Sublayer 150 μm,


top layer <1 μm top layer <1 μm

Effective pore 0.05-10 μm 1-100 nm <2 nm <2 nm


size

Driving force Pressure (<2 bar) Pressure (1-10 Pressure (10-25 Pressure (12-25
bar) bar) bar brackish
water, 40-80 bar
seawater)

Separation Sieving mechanism Sieving Solution-diffusion Solution-diffusion


principle mechanism

Membrane Polymeric, ceramic Polymeric, Polyamide Cellulose


material ceramic triacetate,
aromatic
polyamide,
polyamide and
poly(ether urea)

Applications Water treatment, Dairy, food, Desalination, Desalination of


sterilisation, beverage metallurgy, micropollutant brackish water
clarification, analytical textile, removal, water and seawater,
applications pharmaceutical, softening, production of
automotive, water wastewater, dye ultrapure water,
treatment retention concentration of
juice, sugars and
milk.

Of the membrane filtration processes, MF most closely resembles conventional

coarse filtration, and is suitable for retaining suspensions and emulsions [2]. The

primary rejection mechanism for MF is physical sieving. In addition, the deposit

(cake) on the membrane acts to retain additional smaller particles than would be

expected for a given pore size [3].

9
UF is commonly used to remove colloidal particles and macromolecules. The

main rejection mechanism is physical sieving, although adsorption, charge

interactions, and the self-rejecting action of the deposited layer are also of

importance [3].

RO processes achieve membrane flux and separation by applying high

transmembrane pressure (TMP) to overcome the osmotic pressure of the solution.

RO removes both monovalent and multivalent ions, and is used primarily in

desalination applications. Rejection is largely due to the chemical affinity of the

solute to the membrane material, and typically increases with increasing hydrated

radius of the solute ion. However, interactions between mixtures of ions are

complex, and removal mechanisms are generally poorly understood [3].

NF processes lie between UF and RO processes, with both charge and size

playing important roles in NF rejection. For uncharged molecules, separation

occurs predominantly through physical sieving, while for ionic species, both

sieving effects and electrostatic interactions occur [4].

There are four main types of membrane module [5] which are:

x plate and frame modules, which consist of a thin channel in which one or

both of the walls are a membrane;

x spiral-wound modules, in which a thin channel is wrapped in a spiral

formation to increase membrane area per unit volume of filter;

10
x tubular and hollow fibre modules, which are cylindrical membrane tubes;

and

x capillary fibre modules, which are identical to hollow fibre modules but

with very small inner diameter.

2.1.1 Membrane performance

To describe the behaviour of a membrane filtration system, it is useful to first

define several parameters. The first of these is the flux, or flow though the

membrane, J, which is given by:

1 dV (2.1)
J
A dt

where V is the filtrate volume, t is time, and A is the surface area of the membrane

[3].

The rejection of a membrane Ri describes the efficiency with which the membrane

is able to separate phases of the feed mixture. The rejection is given by:

CP CB  CP (2.2)
Ri 1
CB CB

where CB is the concentration in the bulk of the feed solution, and CP is the

concentration of the permeate. The value of Ri varies between zero (where both

solute and solvent pass through the membrane freely) and one (where the solute is

completely retained).

11
The flow through the membrane can also be characterised by the hydrodynamic

permeability LV, which is given by:

J (2.3)
LV
'P

where ¨3 is the TMP. These terms will be used in the following to describe

theoretical models of the membrane filtration process.

2.2 Membrane flux decline

The primary technical issue which has limited the application of membrane

filtration is that of flux decline –during a membrane separation process, the flux

through the membrane will inevitably decrease with time. The flux decline can be

substantial: in some processes: the process flux is often less than 5% of the pure

water flux [2]. Eliminating or reducing the extent of flux decline is thus one of the

primary aims of membrane research.

This flux decline can be attributed to two main factors: polarisation phenomena

(mainly concentration polarisation), and membrane fouling. A typical pattern of

flux decline over the duration of a membrane filtration run illustrating the

influence of these two factors is shown in Figure 2.2.

12
initial flux

flux decline
due to
polarisation
membrane
flux
flux J decline
due to
fouling

time t

Figure 2.2 - Typical flux decline during filtration run [after 2]

Flux decline due to polarisation occurs relatively rapidly, but is a reversible

process which represents only a small amount of the total flux decline. The

remainder of the flux decline is due to membrane fouling. However, fouling is not

completely independent of polarisation phenomena, and can be minimised by

restricting the extent of polarisation. It is useful to begin by distinguishing

between concentration polarisation (the polarisation phenomenon of interest in

pressure-driven membrane processes) and membrane fouling.

2.2.1 Concentration polarisation

Concentration polarisation occurs when rejected solutes accumulate next to the

membrane wall, forming a boundary layer [6]. This accumulation of solute causes

a diffusive flow back to the feed, which eventually reaches a steady-state

condition where the solute flow to the membrane surface equals the solute flux

through the membrane plus the diffusive flow from the membrane surface back to

13
the feed. This establishes a concentration profile within the boundary layer as

shown in Figure 2.3.

bulk of feed boundary membrane


layer

CW

CB

CP

boundary layer thickness į

distance from membrane surface x

Figure 2.3 - Concentration polarisation profile under steady-state conditions [after 2]

The primary consequence of concentration polarisation is the reduction in

permeate flux, by either increasing the osmotic pressure on the feed side of the

membrane or forming a gel layer at the membrane surface. Several models have

been developed to account for these phenomena, as discussed in the following.

2.2.1.1 Models of concentration polarisation

The Film Theory Model assumes that axial (crossflow) solute convection near the

membrane surface is negligible [7], and describes the concentration profile by a

one-dimensional mass balance equation. By integrating from the membrane wall

to the edge of the boundary later, the following is obtained:

14
DS CW  C P (2.4)
J ln
G C B  C P

where DS is the solute diffusivity, į is the boundary layer thickness, and CW is the

concentration at the membrane surface or wall.

The Gel Polarisation Model assumes that beyond a certain value of pressure

applied to the membrane, the permeation rate is limited by the presence of a gel

layer deposited on the membrane surface, formed as the concentration in the

boundary layer reaches the solubility limit [8]. Thus, assuming an upper limit on

the solute concentration at the membrane surface CG (the gel concentration), the

following equation can be obtained:

DS CG (2.5)
J ln
G CB

The Osmotic Pressure Model assumes that permeate flux is limited by the osmotic

pressure exerted by retained macromolecules on the surface of the membrane [2].

This osmotic pressure is normally neglected as it is small compared to the osmotic

pressure of low molecular weight solutes that pass through the membrane.

However, for high flux and rejection, the high concentration of macromolecules

on the membrane surface means the osmotic pressure can no longer be neglected,

and forms an osmotic pressure difference ¨Ȇ across the membrane. The flux is

then given by:

15
'P  '3 (2.6)
J
KRM

where Ș is the dynamic viscosity of the solvent and RM is the clean membrane

resistance.

The Boundary Layer Resistance model assumes that the high concentration of

retained molecules in the boundary layer exerts a hydraulic resistance upon the

permeating solvent molecules. Assuming this resistance and the hydraulic

resistance of the membrane act in series, the following equation is obtained [9]:

'P (2.7)
J
K R M  RCP

where RCP is the resistance exerted in the boundary layer due to concentration

polarisation. In RO applications, where osmotic pressure is significant, the

osmotic pressure effects are often incorporated into the resistance RCP.

2.2.2 Membrane fouling

Distinct from concentration polarisation, membrane fouling is defined as the

irreversible deposition of retained particles, colloids, emulsions, suspensions,

macromolecules, salts or other substances on or within the membrane [2].

Membrane fouling depends on an array of physical and chemical parameters (such

as concentration, temperature, pH, ionic strength and others) as well as other

specific interactions (e.g. Van der Waals forces, hydrogen bonding, dipole

16
interactions). As such, the phenomenon of fouling is extremely complex and not

fully understood.

Attempts have been made to classify fouling types by several methods, including:

x on the basis of fouling material;

x on the basis of the site at which fouling occurs; and

x on the basis of the underlying mechanism by which fouling occurs.

A brief summary of these classifications follows, though it is emphasised that

there are undoubtedly overlaps and linkages between the different classification

schemes.

2.2.2.1 Fouling materials

Several types of fouling material have been identified to occur in membrane

filtration systems. These include inorganic fouling or scaling, particulate and

colloidal fouling, organic fouling, and biological fouling (biofouling) [10].

Inorganic fouling is due to the deposition of inorganic compounds or scales of the

membrane surface [11]. This occurs when salts rejected by the membrane exceed

their solubility limit in the boundary layer, and are precipitated on the surface.

The inorganic foulants generally of practical concern are salts of calcium,

magnesium, carbonate, sulphate, silica and iron [3].

17
Particulate or colloidal fouling is the fouling of the membrane surface by

particulate material (suspended solids, colloids, and microorganisms) [12].

Particulate fouling is one of the most persistent types of fouling, and is

particularly prominent in the treatment of surface waters.

Organic fouling refers to the accumulation of organic compounds on or within the

membrane, and depends to a large extent on the organic characteristics. Large

organic molecules tend to cause more severe fouling, and mixtures of organic

compounds tend to display worse fouling than in isolation, due to solute-solute

interactions [3].

Biofouling occurs when the membrane surface is exposed to a liquid containing

microorganisms and dissolved organic matter (DOM). The organic molecules are

absorbed to the membrane surface, to which the microorganisms adhere to

produce a biofilm [10]. Biofouling is of significant concern in RO processes,

particularly those in tropical or sup-tropical regions [13].

2.2.2.2 Fouling locations

The classification of membrane fouling by the site of its occurrence distinguishes

between two primary fouling locations: the external surface (the build up of a

cake layer on the feed side of the membrane), and the membrane pores [14].

Fouling within the membrane pores can be further classified as complete pore

blocking (where pores are blocked by particles of approximately the pore size),

18
standard pore blocking (where pores are gradually narrowed by particles much

smaller than the pore size), and incomplete pore blocking (an intermediary

between the two). Internal fouling within the membrane pores is largely a function

of the physico-chemical parameters of the membrane system, while external

fouling is to a larger extent dictated by the hydrodynamic conditions (similar to

concentration polarisation).

2.2.2.3 Fouling mechanisms

The main mechanisms considered to govern membrane fouling are pore blocking

and cake formation (which are related to the site of fouling as discussed above).

Other mechanisms such as adsorption, particle deposition within pores, and

alterations to the cake layer are thought to essentially act by modifying either or

both of the major mechanisms [15].

A further distinction can be made between particle adsorption and deposition.

Macrosolute or particle adsorption is the interactions between particles and the

membrane which occur even in the absence of filtration (i.e. in the absence of a

TMP), and is usually responsible for irreversible fouling. Particle deposition, on

the other hand, is induced by filtration, and is the accumulation of particles on the

top surface of the membrane. This is reversible, non-adhesive fouling which is

analogous to external fouling or cake formation [15].

19
2.2.2.4 Models of membrane fouling

A multitude of mathematical models of membrane fouling have been developed.

A brief summary of the more notable models found in the literature follows

below.

The Resistance in Series Model describes the flux of a fouled MF membrane, and

is an extension of the Boundary Layer Resistance model previously discussed for

concentration polarisation. It adds two additional resistance terms for internal

fouling and external fouling (cake formation), denoted as RP and RC, respectively.

Again assuming the resistances act in series, the flux through the membrane is

then given by:

'P (2.8)
J
K R M  RCP  RC  R P

A set of so-called filtration laws have been proposed [3] to further describe

fouling mechanisms. The general form known as the constant pressure filtration

law is given by:

d 2t § dt ·
n (2.9)
k¨ ¸
dV 2 © dV ¹

where k is a distribution coefficient, and n is a non-negative exponent. This was

used to develop four main filtration models, which are summarised in Table 2.3.

20
Table 2.3 - Summary of models derived for constant pressure filtration law for fouled membranes [3]

Name of model Governing equations Details

Complete Blocking 2
(2.10) Valid for particles with similar size to
d 2t § dt ·
Model k¨ ¸ membrane pores – particles seal the
dV 2 © dV ¹ pores and do not deposit on one
another. J0 is the initial flux through
which can be solved to give
the membrane.

V
J 0 1  e  kt (2.11)

Standard Blocking 3/ 2
(2.12) Valid for particles much smaller than
d 2t § dt ·
Model k¨ ¸ membrane pores – particles pass
dV 2 © dV ¹ through pores and deposit on inner
pore surface. Pore volume decreases
which can be solved to give
proportionally with filtrate volume.

t kt 1 (2.13)

V 2 J0

Intermediate d 2t (2.14) Every particle reaching a pore


§ dt ·
Blocking Model k¨ ¸ contributes to blockage – particles
dV 2 © dV ¹ deposit on one another. Describes
which can be solved to give long term adsorption.

kV ln 1  ktJ 0 (2.15)

Cake Filtration (2.16) Valid for particles much larger than


d 2t
Model k membrane pores – particles deposit on
dV 2 surface without entering pores.
which can be solved to give

t kV 1 (2.17)

V 2 J0

A separate class of models has also been developed for UF processes, in which

mechanisms such as adsorption and charge interactions are more significant. A

summary of the major models for UF processes are summarised in Table 2.4.

21
Table 2.4 - Summary of UF filtration models for fouled membranes [compiled from 3]

Name of model Governing equations Details

Mechanical Sieving Ri >O 2  O @2 for O  1 (2.18) Describes hindered transport of solute


Model by convection and limited by steric
effects. Does not account for solute
Ri 1 for O t 1 (2.19) velocity drag, diffusion, or
concentration effects at membrane
surface.
O
d solute (2.20)
d pore

Modified Sieving (2.21) As per Mechanical Sieving Model,


d solute  2N 1
Rejection Model O except Ȝ includes Debye length term ț-
d pore  2N 1 1
to account for double layer thickness
surrounding charged solute.

Pore Flow Model (2.22) JS is the solute flux, rP is the pore


S'PWn p rp 4 radius, nP is the number of pores, ¨[ is
J
8K'x the membrane thickness and IJis the
tortuosity coefficient. Describes flow
Js 1  V JCB (2.23) through cylindrical membrane pores –
simplified model does not consider
frictional or diffusive forces.

Models for fouled NF processes vary again - the above UF models (and some RO

models to be discussed later) are applicable to NF processes to some extent,

though charge appears to be more significant than for other pressure-driven

processes [3]. NF behaviour for charged molecules can be described by the

extended Nernst-Plank equation [16], which is given by:

dCi ziCi Di , p d< (2.24)


Ji  Di , p  F  K i ,cCiV
dx RT dx

where Ji is the flux of an ion i, Di,p is the ion diffusivity in the membrane, R is the

gas constant, F LVWKH)DUDGD\FRQVWDQWȌLVWKHHOHFWULFDOSRWHQWLDODQGKi,c is the

22
convective hindrance factor in the membrane. The most common model in the

literature, the Donnan steric partioning model (DSPM), first put forward by

Bowen and Mukhtar [17], is based upon Eq. (2.24). In the DSPM, rejection occurs

via both steric effects (when the molecule is physically excluded from the pore)

and by electrical (charge) effects, and the membrane is characterised by three

parameters: the membrane pore radius, the effective thickness/porosity of the

membrane, and the electrical charge inherent to the membrane. The DSPM is

described in detail later in Section 5.3.3. Further extensions to the DSPM have

also been made with the incorporation of dielectric exclusion effects [18, 19], by

which the dielectric constant within the membrane pores is less than that of the

bulk solution, though the details of these approaches are too involved to list here.

Finally, RO filtration models have been developed, which explicitly take into

account the effect of osmotic pressure. Notable models are summarised in Table

2.5.

23
Table 2.5 - Summary of RO filtration models [compiled from 3]

Name of model Governing equations Details

Preferential J B 'P  '3 (2.25) ¨ȆLVWKHRVPRWLFSUHVVXUHGLIIHUHQFH


Sorption / across the membrane, and B is the
Capillary Flow pure water permeability of the
Model membrane. Based on the assumption
of a sorbed water layer at the
membrane surface, which controls
transport through the membrane.

Irreversible J LV 'P  V'3W (2.26) LV is the hydrodynamic permeability


Thermodynamics of the membrane, LS¨ȆW is the
Model osmotic pressure difference between
the membrane wall and permeate, JS is
JS LS '3  1  V JCm.S (2.27) the solute flux, Cm,S is the average
solute concentration across the
membrane and ı is a reflection
'3 (2.28)
V{ for J 0 coefficient. Assumes coupled
'3 transport of solute and solvent.

Solution Diffusion DCm,W Vm,W (2.29) Cm,W is the concentration of water in


Model J 'P  '3 the membrane, Vm,W is the partial
RT'x molar volume of water, ¨[ is the
membrane thickness, and k is a
kDM (2.30) distribution coefficient. Assumes that
J 'CS
'x solute and solvent dissolve in
membrane, and rejection of solute
depends on its ability to diffuse
through structured water inside the
membrane.

It is noted that these models generally consider only a single filtration mechanism

to be active at a given time, when in reality a multitude of separate models may be

required to adequately describe the overall behaviour of the membrane filter. In

addition, most models are only applicable to a single type of membrane process

(MF, UF, NF, RO). There is little work in the literature regarding coupling of

separate models to form a unified model [3].

24
2.3 Methods to reduce flux decline

A multitude of methods have been developed to attempt to reduce the flux decline

phenomenon, though the complex nature of fouling makes a general approach

difficult. The strategies employed fall largely into four groups [15]:

x pre-treatment of the membrane feed solution;

x modification of the membrane properties (surface and materials);

x control of membrane operating parameters; and

x membrane cleaning.

These methods are discussed in the following sections.

2.3.1 Pretreatment of feed solution

The pretreatment of feed solution to the membrane is an effective and well-

studied method of fouling control. Various pre-treatment methods have been used,

including:

x chlorination [e.g. 13, 20];

x coagulation [e.g. 21, 22, 23];

x pH adjustment [e.g. 21, 24, 25];

x activated carbon adsorption [e.g. 26, 27];

x chemical clarification [e.g. 28, 29];

x addition of complexing agents [e.g. 30]; and

25
x use of UF and MF as pre-treatment for RO processes [e.g. 31, 32, 33].

Pre-treatment methods have been extensively covered in the literature on

membrane fouling. For this reason, pre-treatment methods will not be studied in

this thesis. Thorough summaries of the topic may be found elsewhere [e.g. 15].

2.3.2 Modification of membrane properties

The properties of a membrane significantly influence fouling. Mulder [2] notes

that fouling is generally much more severe in porous membranes (UF, MF) than

dense membranes (RO), and that narrow pore size distributions may also reduce

fouling. Attempts have also been made to alter the surface chemistry of the

membrane to discourage fouling as described by Ma et al. [34], including:

x coating of water soluble polymers or charged surfactants onto the

membrane surface;

x forming ultrathin films on the membrane surface using Langmuir-Blodgett

techniques;

x coating of hydrophilic polymers onto the membrane surface; and

x grafting of monomers onto the membrane by electron beam or ultraviolet

radiation.

Details of these processes are beyond the scope of this discussion. An excellent

summary can be found in Hilal et al. [15]. Because of the considerable work in the

literature on this topic and the complexity of membrane materials science and

26
surface chemistry, not to mention the specialised apparatus required for

experimental study, membrane modification methods will not be considered in

this thesis.

2.3.3 Control of membrane operating parameters

Substantial reductions in membrane fouling and flux decline can be made by

manipulation of the operating parameters of the membrane process. This is

typically done with the aim of minimising the effect of concentration polarisation.

This can be achieved by either reducing membrane flux (which is usually

undesirable), or by increasing the mass transfer coefficient k of the system,

defined as:

DS (2.31)
k
G

The mass transfer coefficient characterises the hydrodynamics of the membrane

filtration system, i.e. the value of k is governed by the fluid properties (viscosity,

density, diffusivity), the fluid velocity, and the membrane geometry [2]. The mass

transfer coefficient can be related to some well known dimensionless numbers as

follows:

d (2.32)
kd h b c§
d ·
Sh a Re Sc ¨ h ¸
DS © L¹

27
dhv (2.33)
Re
Q

Q (2.34)
Sc
DS

where Sh is the Sherwood number, Re is the Reynolds number, Sc is the Schmidt

number, a,b,c and d are constants, dh is the hydraulic diameter, L is the length of

the channel, v is the flow velocity, and Ȟis the kinematic viscosity. Generally, an

increase in Reynolds number leads to an increase in the mass transfer coefficient,

due to increased local shear stress on the membrane surface which reduces

concentration polarisation and cake formation. The majority of methods to reduce

concentration polarisation thus focus on promoting turbulent conditions to

maximise shear stress and hence mass transfer. A brief summary of these methods

follows below.

2.3.3.1 Turbulence promoters

Turbulence promoters, or spacers, are a common method of improving membrane

mass transfer. However, the term is slightly misleading, as these devices do not

necessarily induce fully turbulent flow [35]. Spacers work by inducing increased

frictional and pressure losses by increasing resistance to axial flow, as well as

forming localised areas of stagnation (‘dead zones’). Spacers may be

manufactured in several forms, but generally consist of a net-like arrangements of

filaments, where flow disturbances are introduced at the filaments which cross the

28
axial flow direction (either transversely or at some angle). The spacer net can

either be woven, or non-woven, and are typically employed in spiral-wound

membrane modules, as shown in Figure 2.4.

To date, research into spacer design and optimisation has been quite extensive.

Numerous studies have been performed both experimentally [e.g. 36, 37-41] and

computationally [e.g. 42, 43-46] on this topic. It is considered by some authors

[35] that current spacer designs are so well optimised that little progress remains

to be made in this field. While this thesis will not specifically consider spacer

design, modelling of the flow through spacer-filled membrane modules will be

performed as part of developing a general model of the membrane filtration

process.

29
(a)

(b)

Figure 2.4 - Typical net-type spacer geometry, showing (a) use in spiral-wound membrane [47], and (b)

woven and non-woven spacers [35].

2.3.3.2 Gas sparging and bubbling

The injection of air or gases into the membrane feed solution has been

demonstrated by several authors to give significant improvements in membrane

flux [e.g. 48, 49, 50]. A range of flow regimes exists for such two-phase flows: at

high gas flow rates, mist flow or annular flow regimes are observed, and for low

30
gas flow rates, bubble flow is observed [51]. Between these two extremes lie

churning flow and slug flow, which is characterised by a quasi-periodic series of

elongated bullet-shaped bubbles known as Taylor bubbles [52]. The different flow

regimes are illustrated in Figure 2.5.

Figure 2.5 - Flow regimes for gas-injected flow through membrane [35]

The flux enhancement due to gas sparging has been attributed to several

mechanisms, primarily physical displacement of the concentration polarisation

boundary layer, and a reduction in the hydrodynamic resistance of the boundary

layer [49]. However, other significant mechanisms include pressure pulsing

caused by gas slugs, and bubble-induced secondary flow (i.e. flow around the

bubbles, which causes turbulence in the bubble wake) [53]. A significant number

of hydrodynamic and computational studies have been performed on this topic

[e.g. 51, 52, 54, 55, 56], though nearly all have focused exclusively on UF

processes. (These studies will be discussed in greater detail in Chapter 3). While

31
these techniques will not be specifically modelled in the work in this thesis, it

represents a feasible extension of the models developed herein for future work.

2.3.3.3 Dean vortices

Similar to the turbulence promoters discussed previously, Dean vortices have

been suggested as another method for increasing flux by minimising concentration

polarisation. Dean vortices are secondary flows which occur in flow through

coiled pipes due to centrifugal forcing. However, they may also be formed by use

of helical screw thread shaped inserts in tubular membranes, as shown in Figure

2.6. Bellhouse et al. [57] demonstrated substantial improvements in membrane

flux using these inserts both experimentally for UF, MF, and NF processes, and

using computational methods to predict the formation of the Dean vortices.

Figure 2.6 - Helical screw thread inserts used to induce Dean vortices (left: one start thread, right: three start

thread) [57]

Other authors have also demonstrated that the action of Dean vortices in curved

membrane tubes (typically helically or sinusoidally curved) is even more effective

32
in reducing the effects of concentration polarisation [e.g. 58, 59-61]. However, at

present such membrane geometries are not commercially available. Again, while

not specifically considered in this thesis, this remains another area that could in

future be investigated through use of the general computational models to be

developed in this thesis.

2.3.3.4 Pulsatile flow

The use of pulsatile flows (feed flows varying in velocity or pressure with time)

has also been demonstrated to enhance permeate flux. A variety of methods have

been used to generate pulsatile flows, including:

x collapsible-tube oscillators [62];

x infrasonic pulse generation [63];

x insertion of rotating perforated discs into the entrance of tubular

membranes [64]; and

x pneumatically activated jack / piston arrangements [65].

A number of mechanisms have been suggested for the flux increase afforded by

pulsatile flow, including increased shear, increased radial transport of particles

away from the membrane surface, backflushing though pores, cavitation, and

membrane flexing [62]. The shape of the pulsatile waveform has been observed to

significantly affect the efficiency of the process, with negative pressures (i.e.

backflushing) necessary to achieve the greatest reduction in membrane fouling

33
[66]. While a fair amount of research into pulsatile flows has occurred, the

majority of work has been experimental. There is some potential for further work

in this area using computational methods to further characterise and optimise the

mechanisms of flux enhancement by pulsation, though this is not specifically

modelled in this work.

2.3.3.5 Ultrasonic methods

The passing of ultrasonic waves through the filtration solution has been suggested

as a means for minimising membrane flux decline. As ultrasonic waves of

sufficiently high power are propagated through a liquid, cavitation bubbles can

occur, typically at frequencies of 20-1000 kHz [67]. Each cavitation bubble can

generate local temperatures of 4000-6000K and pressures of 100-200 MPa. The

inrush of fluid to fill the collapsed bubble forms a targeted jet which can be used

to increase shear stress at the membrane surface and hence minimise

concentration polarisation. A range of other potential mechanisms for particle

removal or detachment have been suggested, including cavitational phenomena

(microstreaming), acoustic streaming, and micro-jetting, which are shown in

Figure 2.7.

Most applications of ultrasonics to membrane filtration have focused on

membrane cleaning after filtration has occurred [e.g. 68, 69-71]. However, several

investigations have also been made into using ultrasonics during the filtration

process to reduce fouling with encouraging results [e.g. 72, 73]. However, there

34
has been experimental evidence of damage to the membrane surface by ultrasonic

treatment [e.g. 74, 75] . Overall, the use of ultrasonics in membrane filtration is

relatively unstudied and could be a potential application of a computational

membrane filtration model, though not considered further in this thesis.

Figure 2.7 - Particle detachment mechanisms for ultrasonic agitation [68]

2.3.4 Membrane cleaning

The methods outlined above are generally concerned with preventing or

mitigating the effects of flux decline via minimisation of concentration

35
polarisation and membrane fouling. Membrane cleaning methods, on the other

hand, are essentially a remedial method for dealing with the after-effects of

membrane fouling. It is considered in this case that the adage ‘prevention is better

than cure’ is fairly accurate; hence, this study will not consider membrane

cleaning methods in any great detail. Concise summaries of the common

membrane cleaning methods can be found elsewhere [e.g. 11, 76].

2.4 Summary and directions for research

It is apparent that considerable research effort has been invested on describing the

effects of various parameters on membrane flux decline. However, there is still

substantial work required to gain a full understanding of the fundamental

mechanisms of flux decline, including concentration polarisation and membrane

fouling. This is largely due to the complexity and variability of membrane

processes, and the difficulty in using analytical mathematical models to describe

these systems. The use of computational methods for simulating the mechanisms

of flux decline therefore represents an important direction for future study.

Computational simulations of the membrane filtration process also offer a

relatively inexpensive option for developing and optimising methods to reduce the

incidence of flux decline. A number of methods have been suggested or are

currently practiced, including:

x feed pretreatment;

36
x modification of membrane properties;

x membrane cleaning methods;

x control of membrane process operating parameters.

The first three classes of methods will not be considered in this study, as they

have generally been well-studied, are not always applicable to the general

problem of membrane flux decline, or are not suited to computational simulation.

However, control of process operating parameters appears to be a potentially

workable solution for minimising membrane flux decline. A valuable method for

investigating the control of these parameters is the use of computational

simulation using computation fluid dynamics techniques, which are discussed in

detail in the following chapter.

37
Chapter 3

The use of computational fluid dynamics in membrane

filtration

Computational fluid dynamics (CFD) refers to the use of numerical methods with

digital computers to solve fluid flow problems. The basic premise of CFD is

replacing the partial differential equations which govern viscous fluid flow with

discretised algebraic equations that approximate the partial differential equations

[77]. These equations are then solved numerically to obtain solutions for problems

where analytical solutions to the governing equations are not available.

CFD methods have relatively recently become a viable tool for analysis of

membrane filtration systems, due to increases in computing performance and

developments in numerical and mathematical methods. This chapter gives a brief

overview of the basic principles of CFD and current applications in fluid

dynamics, then discusses in detail various approaches to describing the behaviour

of membrane filtration systems using CFD.

38
3.1 Basic principles of CFD

Viscous fluid flow is governed by the basic principles of conservation of mass,

and conservation of linear momentum, which can be described by the following

equations, respectively:

wU (3.1)
 ’ ˜ Uv 0
wt

w Uv  wP (3.2)
 ’ ˜ Uvu  ’ ˜ P’v  SM y
wt wy

where ȡ is the density of the fluid, v is the component of the fluid velocity in the x

direction, and SMy is a source term in the y direction [35].

To solve the partial differential equations governing fluid flow, appropriate

numerical discretisation schemes must be employed. The three main classes of

methods are finite difference methods (FDM), finite element methods (FEM) and

finite volume methods (FVM). A detailed discussion of these methods can be

found elsewhere [78-80]. The main characteristics of each method are listed in

Table 3.1.

39
Table 3.1- Summary of characteristics of discretisation methods [compiled from 79]

Method Formulation considerations Mesh considerations Boundary condition


considerations

FDM Conceptually easy to Meshes must be Neumann boundary conditions


formulate structured in two or three (derivatives of variables specified
dimensions. Curvilinear at boundaries) cannot be strictly
meshes must be enforced, only approximated.
transformed to structured
Cartesian coordinates.
More difficult to
accommodate complex
geometries.

FEM Formulation less straight Easier to accommodate Neumann boundary conditions


forward complex geometries. are exactly enforced.
Non-Cartesian
coordinates and
unstructured meshes
possible.

FVM Can be formulated based on Depends on formulation Neumann boundary conditions


FEM or FDM (FEM or FDM). are exactly enforced.

Theoretically, it is possible to solve the governing equations of viscous flow via

computational methods for both laminar and turbulent flow regimes. However, in

turbulent flows, there is a much larger range of length and time scales than in

laminar flows, making the equations for turbulent flow more difficult and

computationally expensive to solve [80]. The problem of turbulence can be

approached in three main ways [79]:

x direct numerical simulation (DNS), where refined computational meshes

are used to resolve all scales of turbulence, large and small. This is

generally not feasible for problems of practical interest;

40
x Reynolds averaged Navier-Stokes (RANS) methods, where turbulent

motion is described in terms of time-averaged quantities, so that time-

dependent flow quantities are given in terms of a time-averaged mean

component and a time-varying fluctuating component. All scales of

turbulence are modelled (as opposed to directly computed) using these

methods;

x large eddy simulation (LES), where large scale turbulence is directly

computed and small-scale turbulence is modelled using time-averaged

methods. This method is becoming increasingly popular, as it offers

additional resolution of large-scale turbulence without the heavy

computational cost of DNS methods.

A detailed treatment of individual turbulence models is beyond the scope of this

thesis. Some of the further discussion will refer to certain turbulence models; the

reader is referred elsewhere for details [79, 80].

3.2 Modelling of membrane processes under laminar conditions

3.2.1 Flow through porous membranes

Investigations into the hydrodynamics of membrane channels largely began with

Berman’s [81] study of laminar flow in a rectangular channel with porous walls,

in which expressions were derived for fluid velocity and pressure as functions of

the channel dimensions, fluid properties and position within the channel. The

41
expressions were derived from the Navier-Stokes equations with the following

assumptions: (1) a steady state prevails; (2) the fluid is incompressible; (3) no

external forces act on the fluid; (4) the flow is laminar and (5) the velocity of the

fluid leaving the walls of the channel is independent of position [81]. The primary

results derived were that pressure drop in a porous channel is significantly less

than a non-porous channel, and that the velocity profile observed within the

channel was flatter near the centre and more curved near the walls than the

classical parabolic profile of Poiseuille flow.

Later studies were conducted by Friedman and Gillis [82] for a porous tube with a

viscous fluid, which were further refined by Mizushina et al. [83] for the case of

variable radial mass flux, and by Galowin and De Santis [84] for the case of

variable wall suction [all cited in 35]. Berwin’s approach was also extended by

Karode [85], who assumed that the velocity of fluid exiting the porous wall was

not constant, but was proportional to the local TMP. This revised analytical

solution showed good agreement with Berwin’s results for the case of constant

wall velocity.

Recent numerical treatments of laminar flow through a porous membrane have

followed a similar approach, by using the Navier-Stokes equations to describe the

motion of the fluid through the main ‘channel’ of the membrane module, and the

Darcy equation to describe the fluid flow through the porous walls of the

membrane module. For the case of flow through a cylindrical tube, the continuity

and Navier-Stokes equations can be written in cylindrical coordinates as:

42
wvr vr wvz (3.3)
  0
wr r wz

wvr wv wp w § wv · (3.4)
Uvr  Uv z z   ¨ 2K r ¸
wr wz wr wr © wr ¹
wvr v w ª § wv wv ·º
 2K  2K r2  «K ¨ z  r ¸»  Ug r 0
wr r wz ¬ © wr wz ¹¼

wv z wv wp w § wv z · (3.5)
Uvr  Uv z z   ¨ 2K ¸
wr wz wz wz © wz ¹
K § wv wv · w ª § wvz wvr ·º
 ¨ z r ¸  «K ¨  ¸»  Ug z 0
r © wr wz ¹ wr ¬ © wr wz ¹¼

where r and z represent the radial and axial coordinates, respectively.

Similarly, the Darcy equation can be written as:

Pf wp (3.6)
vr  0
Lr wr

Pf wp (3.7)
vz  0
Lz wz

where μf is the viscosity of the filtrate, and Lr and Lv are the permeability of the

porous wall in the radial and axial directions, respectively.

An example of this approach is the work by Nassehi [86], who presented a

method for coupling the two sets of equations using FEM. The lack of second-

order derivatives in the Darcy equation (which are present in the Navier-Stokes

equations) presents difficulties in coupling the two sets of equations. Hence, the

43
Brinkmann equation which does contain second-order terms (shown below) has

sometimes been used in place of the Darcy equation, though this is not ideal, as

the Brinkmann equation is limited to high porosities generally not found in

membrane modules. The Brinkmann equation is given by:

P (3.8)
'P  P’ 2v  v
K

Nassehi resolved this issue by developing a Galerkin finite element scheme based

on the use of C0 Lagrange elements, in which the layer of elements joining the

porous wall to the rest of the solution domain are cast in such a form to represent

a combined flow field. That is, elements in the stiffness matrices corresponding to

the nodes located on the porous wall are replaced by the discretised form of the

Darcy equation, while the remainder of the matrices remain as per the Navier-

Stokes equations (the reader is referred to Nassehi’s paper for further details of the

finite element scheme). Results demonstrated the scheme was flexible in

application and accurately preserved flow continuity.

Another method of coupling the Navier-Stokes and Darcy equations was

demonstrated by Damak et al. [87] by assuming the permeability of the wall Lv to

be the same in the axial and radial direction, and assuming the viscosity of the

inflow and filtrate to be identical (denoted by Ș). This allows the flow through the

porous wall to be taken simply as a boundary condition for the Navier-Stokes

equations governing the free flow through the tube. The boundary condition can

be written as:

44
Lv 'P (3.9)
vr  , vz 0 at r R, and 0 d z d L
K e

where e is the thickness of the porous wall, and R and L are the radius and length

of the tube, respectively. The justification for setting the axial velocity at the wall

to zero is given by work carried out by Schmitz and Prat [cited in 87, 88], which

demonstrated that the slip velocity at the porous wall surface was negligible. The

Navier-Stokes equations can then be solved by an implicit finite difference

method. The model was verified by comparison against velocity profiles

experimentally measured by Gouverneur [cited in 87, 89], and showed good

agreement. The results also demonstrated that the deviation for the parabolic

Poiseuille profile previously predicted by Berman [81] was directly dependent on

the wall suction (i.e. the local TMP).

3.2.2 Concentration-polarisation effects

The inclusion of concentration polarisation effects into the treatment of porous

walls introduces additional difficulties. The boundary layer formed by

concentration polarisation can be viewed as a kind of second porous wall with a

lower permeability than the membrane. The primary difficulty in modelling this

situation is determining appropriate concentration boundary conditions at the wall,

as concentrations will be continually changing at the wall, and the wall geometry

itself may change over time due to particle deposition. Several approaches have

been developed to account for this, and are discussed below.

45
A relatively simple approach to this problem was used by Huang and Morrissey

[90]. They used FEM to simulate the development of the concentration

polarisation boundary layer in a rectangular crossflow ultrafiltration membrane by

solving the Navier-Stokes equations to determine fluid velocities in a manner

similar to Berman [81]. These were then used to solve the convective-diffusive

mass transfer equation using FEM, which describes the formation of the boundary

layer:

wC v wC D w 2C (3.10)
u 
wx H wO H 2 wO2

with the boundary C C0 at x 0 (3.11)

conditions

wC (3.12)
0 at O 0
wO

wC JW C (3.13)
at O 1
wO D

where JW is defined as per Equation (2.6). The finite element package Pdease2D®

was used, with a fine mesh near the membrane surface to accommodate the large

concentration gradient in the vicinity of the boundary layer, as shown in Figure

3.1. Results from the simulations were compared to experimental results using

bovine serum albumin (BSA) solutions. The results demonstrated a linear

relationship between the diffusion coefficient and the thickness of the

concentration polarisation boundary layer, as predicted by the classical Film

Theory and Gel Polarisation models. The model also corresponded well with

46
experimental predictions for the mass transfer coefficient and concentration on the

membrane surface.

Figure 3.1 - Finite element mesh used to model concentration polarisation phenomena [90]

Richardson and Nassehi [91] used a more sophisticated approach by assuming

that the solution does not remain evenly mixed in domains with porous walls. At

each time step, the physical properties of the fluid (viscosity and density) are

updated using relations for the concentration calculated at each time step, as

shown in Figure 3.2. In their model, the fluid flow in the bulk of the channel is

described by the Navier-Stokes equations, while the mass transfer through the

porous wall is described by the two-dimensional convective dispersion equation,

which is written as:

47
wC wC wC ªw º wC ª w º wC (3.14)
wt
 vx
wx
 vy
wy « wx D C » wx  « wy D C » wy
¬ ¼ ¬ ¼

The solution of the flow and convective dispersion equations is then carried out in

a decoupled fashion (i.e. in separate steps, with the Navier-Stokes equations

solved to provide velocity and pressure fields, with the velocity field then used to

solve the convective dispersion equation). The density and viscosity are then

updated using current values of velocity and concentration, and the process

repeated until the solution converges. The density is calculated in terms of a solids

YROXPH IUDFWLRQ ĭ ZKLFK Fan be determined from the concentration, and the

densities of the solid and fluid (ȡs and ȡf, respectively), as written below:

U=U s )+U f ( 1-)) (3.15)

The viscosity is calculated by a two-step process, where the viscosity is first

updated using the following power law model:

K=K oJ n 1 (3.16)

where Șo is the consistency coefficient, n is the power law index, and J is the

shear rate. The viscosity is then updated based on the solids volume fraction via a

relationship such as the following:

Kf (3.17)
=1  /)
K

48
where Șf LVWKHYLVFRVLW\RIWKHFDUU\LQJIOXLGDQGȁLVDQHPSLULFal factor found

experimentally. The equations were solved by use of a streamline upwind Petrov-

Galerkin FEM technique, details of which can be found elsewhere. Results

indicated the model was capable of predicting both flow and mass transport

patterns for a variety of domains including flat and curved porous walls.

Figure 3.2 - Solution algorithm proposed by Richardson and Nassehi [91]

49
3.2.3 Cake formation effects

Geissler and Werner [92] simulated crossflow filtration in a flat-sheet membrane

system considering the formation of a cake layer upon the membrane. The growth

of the cake layer was modelled by considering both the convective transport of

particles towards the membrane and the diffusive transport in the other direction

due to concentration polarisation. The following assumptions were made in

developing the model:

x the flow is two-dimensional, incompressible, laminar, and Newtonian,

characterised by a local suspension viscosity Șsus;

x the particles within the fluid are neutrally buoyant;

x the permeate contains no particles; and

x the cake layer formed is immobile.

The equations thus derived to describe the particle transport within the system are

as follows:

w § * wC · w § * wC ·
(3.18)
¨ uC  D ¸  ¨¨ vC  D ¸ 0
wx © wx ¹ wy © wy ¸¹

2 wu (3.19)
D* Dˆ Iv rp
wy

'P x (3.20)
vF x
K >RM  Rk x @

50
G k x (3.21)
Rk x
Kk

dG k x
>dm x  dm
k x @dt
dif (3.22)
U p 1  H k dA x

where D* is the shear-induced diffusion coefficient, Dˆ Iv is the dimensionless

diffusion coefficient, rp is the particle radius, u and v are the velocities in the axial

and normal dimensions with respect to the membrane, Rk is the hydraulic

resistance of the cake, vF is the permeate velocity, Kk is the cake permeability, įk

is the local cake height, İk is the cake porosity, m is the mass flux and the

subscripts k and dif represent the cake and diffusion, respectively.

The model was verified against experimental observations of permeate flux and

cake layer height, and showed good agreement (within 15% of predicted values).

However, the model requires empirical parameters (particularly the hydraulic

resistance of the cake layer) to be determined by analysis of experimental data in

order to achieve useful results.

Lee and Clark [93] developed a more advanced model of cake formation for a

rectangular crossflow ultrafiltration system. Again, mass transfer is expressed by

the two-dimensional convective dispersion equation, and the steady-state

concentration field can be obtained by solving the following equations:

51
wC wC w 2C (3.23)
u v Dy
wx wy wy 2

with the boundary C x 0, y CB (3.24)

conditions:

§ wC · (3.25)
¨¨ ¸¸ 0
© wy ¹ y 0

§ wC · (3.26)
¨¨ vC  D y ¸ 0
© wy ¸¹ y H0

where the half height of the channel is given by H0, and Dy is the diffusion

coefficient in the transverse direction. Equation (3.26) can be interpreted as

meaning there is no particle accumulation on the membrane surface (i.e. a steady-

state has been reached), which is valid for a membrane with complete rejection.

The steady-state model is then adapted to provide a step-wise pseudo-steady-state

model of the flux decline due to cake formation. This is done by assuming the

particle concentration at the membrane surface CW does not exceed a limiting

value Cmax U p 1  H k , and that the excess of particles will combine to form a

cake layer at the surface.

The solution algorithm proposed is as follows; first, the concentration profile is

obtained by solving the steady-state model using the velocity field at the previous

time step. Second, it is determined whether flux decline occurs by comparing the

concentration profile with the calculated Cmax – if greater than Cmax, a cake layer

52
will be formed at that location. The cake growth is calculated by a mass balance

between the mass convection rate and the mass accumulation rate at the

membrane surface. The permeate flux decline is then calculated using a

resistance-in-series model similar to Equation (2.8), so the local permeate flux is

known at each axial location. Third, a new velocity field is calculated using the

updated permeate flux as a boundary condition, and concentration profiles are

again calculated using the steady-state model. The second and third steps are then

repeated until CW d Cmax at every axial location. The solution algorithm is

summarised in Figure 3.3.

53
Figure 3.3 - Solution algorithm for Lee and Clark [93] cake formation model

The model results presented by Lee and Clark showed good agreement with

experimental results; however the model still requires the input of the specific

resistance of the cake deposit, which can only be determined by dead-end

filtration experiments.

54
3.2.4 The influence of osmotic pressure

Osmotic pressure plays an important role in RO and NF membranes. Several

authors have attempted to include the effects of osmotic pressure within CFD

models, by incorporating boundary conditions for the porous membrane wall

which can be expressed as a function of the osmotic pressure.

Geraldes et al. [94] developed a numerical model using FVM to simulate NF

processes in both spiral-wound and plate-and-frame systems, which can be

modelled as a slit-type channel configuration (i.e. the height of the channel is very

small compared to the length and width of the channel). These geometries mean

that the effect of the concentration polarisation boundary layer is a significant

operating issue. The model developed by Geraldes et al. [94] uses CFD for the

modelling of the fluid phase, with boundary conditions that take into account the

solute transport occurring within the membrane, including osmotic pressure

effects. The transport equations used for modelling of the fluid were as follows:

w Uu w Uv (3.27
 0
wx wy
)

w Uuu w Uvu wp w § wu · w ª § wu wv ·º 2 w ª § wu wv ·º (3.28


   2 ¨ P ¸  « P ¨¨  ¸»  «P ¨  ¸»
wx wy wx wx © wx ¹ wy ¬ © wy wx ¸¹¼ 3 wx ¬ ¨© wx wy ¸¹¼
)

w Uuv w Uvv wp w § wv · w ª § wu wv ·º 2 w ª § wu wv ·º (3.29


   2 ¨¨ P ¸¸  « P ¨¨  ¸»  «P ¨  ¸»
wx wy wy wy © wy ¹ wx ¬ © wy wx ¸¹¼ 3 wy ¬ ¨© wx wy ¸¹¼

55
)

w UuZ A w UvZ A w ª § wZ ·º w ª § wZ ·º (3.30


 « UD AB ¨ A ¸»  « UD AB ¨¨ A ¸¸»
wx wy wx ¬ © wx ¹¼ wy ¬ © wy ¹¼ )

where ȦA is the mass fraction of species A within the solution, and DAB is the

binary mass-diffusion coefficient. Considering the channel to have length l and

half height h, the boundary conditions are given by:

x 0, y, u u0 , v 0, Z A Z A0 (3.31)

wu wv wZ A (3.32)
x l , y , 0, 0, 0
wx wx wx

wZ A (3.33)
y 0, x, u 0, Uv U p v p , UD AB U p v pZ A f c
wy

wZ A (3.34)
y 2h, x, u 0, v 0, 0
wy

where u0 is the uniform inlet velocity, ȦA0 is the uniform mass fraction at the inlet,

f’ is the intrinsic rejection coefficient and the p subscript denotes properties

relating to the permeate. The intrinsic rejection coefficient f’ can be described by

the Deen hindered transport model, which for the case of Pe << 1 can be

simplified as:

ª § 8u ·§ HD AB
0
·º
1 (3.35)
fc  ¨
« ¨ 2 ¸¨
1 ¸¨ ¸»
¸
¬« © r0 ¹© 'P ¹¼»

56
where r0 is the average pore radius of the membrane, H is a hindrance factor for

solute diffusive transport and D AB


0
is the binary mass-diffusion coefficient at

infinite dilution. The permeate velocity vp is dependent upon the TMP and the

osmotic pressure. However, studies by other researchers [cited in 94, e.g. 95, and

96] have established that permeate velocity declines with increased solute

concentration at the membrane surface (i.e. in the presence of a concentration

polarisation layer). The authors thus introduced a concentration dependent

correction factor ȁ to account for this, in a manner similar to Rosén and Trägårdh

[cited in 94, 97]. The permeate velocity is then given by:

vp  /Z Am Ah 'P  '3 (3.36)

'3
3 Z Am  Z Ap (3.37)

where Ah is the hydraulic permeability of the membrane, and the subscripts m and

p refer to the membrane surface and the permeate, respectively. The above

equations were discretised using a finite volume formulation. Comparison of the

computational results showed excellent agreement with experimental results.

Essentially identical approaches to describe the permeate velocity, albeit without

the use of a correction factor, were also employed by Fletcher and Wiley [98] and

Miranda and Campos [99].

57
3.3 Modelling of membrane processes under turbulent conditions

The studies previously mentioned in this chapter have all concerned laminar flows

within the membrane module. However, it has been widely suggested that

turbulent conditions do occur in practice in UF systems [5, 100]. Further, the

mixing afforded by turbulent flow may provide significant benefits in reducing

concentration polarisation effects. Investigation of turbulent membrane processes

is an obvious application of CFD, though it poses significantly more theoretical

difficulties than the laminar case.

One of the obvious questions relating to the use of CFD to simulate membrane

processes is the question of when turbulent models need to be employed.

Typically, flow in narrow channels remains laminar up to Reynolds numbers of

approximately 2000. However, studies have indicated that this is not the case for

some types of membrane modules; for example, Schwinge et al.[44] examined the

transition to turbulence (i.e. the appearance of unsteady flow structures) in spiral-

wound membrane modules, concluding that transition may occur in spacer-filled

channels at Reynolds numbers of 300 to 400. Ranade and Kumar [101] performed

direct numerical simulations (DNS) on spacer-filled channels, which indicated a

transition to turbulence at a Reynolds number of approximately 350. Belfort and

Nagata [102] also showed the transition to turbulence in tubular channels is

delayed to a Reynolds number of approximately 4000. There is limited

information in the literature for other types of membrane modules.

58
However, the appearance of unsteady flow structures and the transition to

turbulence does not necessarily mean that employing conventional turbulence

models is the best approach. Turbulence models are suitable only for high

Reynolds numbers (generally greater than 30 000), at which point turbulence can

be assumed to be fully developed and isotropic [103]. Pellerin et al. [5] provide

some examples of UF and MF flow with Reynolds numbers of 20 000 and 30 000,

respectively, based on the guidelines of Belfort and Nagata [102]. However, the

majority of membrane flows are below this level; for example, it is reported that

for spiral-wound membrane modules, most flows lie in the range NRe = 1000 to

3000 [e.g. 42, 100]. Given this fact, other methods to describe unsteady flows in

these flow regimes may be more suitable, such as direct numerical simulation

(DNS), and we discuss this subject in greater detail later in Section 4.3.1. We first

begin however with a review of studies using conventional turbulence models.

3.3.1 Studies using Reynolds averaged Navier-Stokes methods

Pellerin et al.[5] proposed a general hydrodynamic model with the ability to

incorporate turbulent effects, as well as complex geometries and pressure-related

boundary conditions (to describe the mass transfer through the membrane

surface).. The transport equation used can be written in general terms as:

w UM
 ’ UuM  *M ’M
(3.38)
SM
wt

59
ZKHUHīij is the exchange coefficient, Sij is the volumetric source rate, and ijis a

general field variable. Turbulent effects in the field variable are represented using

a RANS approach via a combination of a time-averaged component M and a time-

varying component M c :

M M  Mc (3.39)

The two equation k-İ turbulence model was used to resolve the turbulent motion,

where k is the turbulent kinetic energy and İ is the turbulent energy dissipation

rate. Further discussion on the k-İ model can be found in a multitude of sources

[e.g. 5, 80]. Boundary conditions at the membrane wall are given by the

transmembrane velocity vw using the Darcy equation:

vw A 'P  V'3 (3.40)

where A is a membrane-specific coefficient. The reflection coefficient ı is a

measure of the rejection of the membrane, and was set to one (very high

rejection). Finally, the concentration field of the solute can be written in the

generalised transport equation format as:

ª §U ·º (3.41)
w « U ¨¨ k ¸¸»
¬ © U ¹¼ ª §U
 ’ « Uu ¨¨ k
· §U
¸¸  DU’¨¨ k
·º
¸¸» 0
wt ¬ © U ¹ © U ¹¼

where ȡk is the density fraction of species k. The above equations were discretised

in finite volume form using the computer code TURCOM, and simulations were

60
run to represent RO, UF and MF conditions. The results showed good agreement

with classical analytical solutions [5, 81, 104] for low Reynolds numbers.

However, significant deviation from the velocity profiles in the analytical

solutions was observed for high Reynolds number cases (as high as 50 000), as

shown in Figure 3.4.

(a) (b)

Figure 3.4 - Deviation of analytical and numerical velocity profiles for Pellerin et al.'s model for (a)

ultrafiltration, (b) microfiltration [5].

Ranade and Kumar [43] also carried out simulations using the k-İ turbulence

model to investigate spacer effects in spiral-wound membranes. The turbulence

model was only employed for Reynolds numbers greater than 300, based on

recommendations from the literature and the authors’ own direct numerical

simulations. The commercial package FLUENT was used for all simulations. The

61
authors successfully managed to model a variety of spacer geometries, with

adequate agreement with experimental results reported by Da Costa et al. [cited in

43, 105, 106]. They also noted that full CFD simulations of an entire membrane

module under turbulent conditions remain very computationally expensive, but

good results could be obtained by considering a representative unit ‘cell’ of the

module.

Cao et al. [107] also employed a RANS approach in simulating turbulence effects

caused by net-type spacers in a narrow channel. The authors made use of the RNG

k-İ turbulence model, which differs slightly in formulation from the classic k-İ

model used by Pellerin et al. [5]. The RNG model aims to describe the effects of

small eddies in terms of larger motions and a modified viscosity term in order to

eliminate these small-scale eddies from the governing equations. The RNG model

is particularly effective when small-scale flows are significantly anisotropic, as

such anisotropic flows tend to either become decoupled from the large-scale flow

or give their energy to large-scale eddies, rather than interacting directly with the

large-scale flow by increasing the dissipation of large eddies, as was demonstrated

by Sivashinsky and Yakhot [108]. The region adjacent to the membrane surface is

typically highly anisotropic, making the RNG approach suitable to this

application. Simulations were carried out with the FLUENT package for a variety

of geometric spacer configurations, details of which are not necessary to be

presented here.

62
3.3.2 Studies using direct numerical simulation methods

Direct numerical simulation (DNS) methods allow small-scale and large-scale

turbulent effects to be calculated directly, without resorting to the approximation

of turbulence models. This is considered to be a more useful approach for most

membrane filtration conditions, where Reynolds numbers are sufficiently low

(generally less than 30 000) that turbulence is not fully developed, and

conventional turbulence models are not strictly applicable. However, this

approach is extremely computationally expensive for most problems of practical

interest. Nonetheless, several studies have been undertaken using these methods.

Miyake et al. [109] made use of a DNS approach, where all scales of turbulence

are directly computed, to simulate turbulent flow with a periodic pressure gradient

in a rectangular channel with one porous wall. Using spectral methods developed

by Kim et al. [cited in 109, 110] to solve the three-dimensional Navier-Stokes

equations, a flow was simulated with periodic pressure boundary conditions on

the porous wall. The study indicated that for this flow, good results (compared to

the direct numerical solution) could be obtained by large-eddy simulation (LES)

models, as the periodic pressure gradient has less influence on small-scale eddies

compared to large eddies. This result is interesting, given the relative scarcity of

LES models in the membrane modelling literature. However, the applicability of

this result to more realistic membrane operating conditions is uncertain.

63
More recently, several investigations have been made using commercial CFD

codes to perform transient simulations with very small timesteps and fine

computational meshes to resolve unsteady flows in membrane channels. Some

examples of this approach can be seen in studies by Schwinge et al. [44, 111, 112]

and Fimbres-Weihs et al. [113, 114] using the ANSYS CFX code, and Koutsou et

al. [42] using FLUENT. These models were used to study the effect of spacers in

rectangular membrane channels, with all cases considered in these studies

simplified by the assumption of impermeable walls or the use of a ‘dissolving-

wall’ boundary condition, where the wall concentration is set to a high fixed value

to simulate the effects of concentration polarisation. Koutsou et al [42] found that

instabilities in the vicinity of the cylinder began to occur at a Reynolds number of

60, and further instabilities in the form of wall eddies began to occur at a

Reynolds number of 78. Schwinge et al.[44] also identified transition to unsteady

flow occurring at Reynolds numbers between 300 and 600, dependent on the

geometrical configuration of the spacers. These models provide useful insight into

the unsteady flow behaviour of these systems, but the simplified wall boundary

conditions employed are not entirely physically representative of a real membrane

system.

Newer studies have incorporated more realistic descriptions of membrane channel

permeation, such as that of Wardeh and Morvan [115]. This study simulated

unsteady flow in 2D spacer-filled channels for conditions similar to that of [44],

but incorporated a permeable wall condition using a Darcy’s law type expression.

64
It appears reasonable that further progress in this area should inevitably include

the effects of membrane permeation in this manner to develop a more fully

representative hydrodynamic model.

Finally, the simulation of flow in spacer-filled channels for these conditions has

also led to the use of 3D models becoming somewhat common. While 3D models

provide a more accurate representation of real flow conditions, they also incur

greatly increased computational expense. Consequently, initial studies used low

spatial resolutions [116] or neglected the membrane mass transport processes by

assuming impermeable walls [e.g. 101, 116, 117]. The most important and

common method of reducing computational expense for these 3D simulations

however is through use of a unit cell approach, using spatially periodic or

‘wrapped’ boundary conditions. This exploits the fact that spatially periodic (but

not necessarily temporally periodic) flow structures tend to occur in spacer-filled

channels between filaments, and hence simulation of only a single unit cell

between repeating filament structures is required. There are multiple examples of

this approach using the dissolving-wall boundary condition to ease computational

requirements [40, 118-121], with one study by Lau et al. [122] employing a true

permeable wall boundary condition. These studies have also almost universally

used the approach of direct numerical simulation of unsteady flows by use of

laminar models with very fine spatial and temporal discretisation. However, due

to the expense of 3D simulations, this approach is still limited to the unit-cell

approach, where spatially periodic flow structures occur. In this respect, direct

65
numerical simulation in this fashion remains more feasible in a general case by

use of 2D models.

3.3.3 Summary of turbulent membrane models

It is clear that nearly the entirety of research in the literature related to turbulence

modelling in membrane modules focuses on the modelling and optimisation of

spacers and turbulence promoters in spiral-wound membranes. In addition, most

applications of turbulence models have employed the standard k-İ model, or the

RNG k-İ model, which are suited only to high Reynolds numbers which may not

be common to a wide range of membrane applications. Little work has been done

in applying turbulent CFD techniques to other membrane configurations, and

correspondingly little work has been done using LES techniques, even though

these are becoming increasingly popular within the discipline of CFD at large.

However, the use of transient laminar flow simulations with very fine meshes and

timestepping has also been shown to well describe unsteady flow behaviour for

lower Reynolds numbers, which is where the majority of membrane applications

are likely to occur.

3.4 Prediction of membrane mass transfer

The majority of the CFD models discussed so far in this chapter have generally

attempted to model only the hydrodynamic conditions within the membrane

channel or module in question. Most neglect what is the most important property

66
of the membrane, which is the ability to selectively remove constituents from a

feed stream; that is, the transfer of mass through the membrane itself. A variety of

models have been developed to describe this mass transfer process, as outlined in

Section 2.2.2.4. By coupling a predictive mass transfer model to a CFD model, a

reasonably general model of the membrane filtration process would be potentially

realisable.

However, due to the different removal mechanisms in the different types of

pressure-driven membrane filtration, it is difficult for a single model to describe

membrane mass transfer generally. For the lower-pressure processes, where

particulate constituents in the feed stream are common and the membrane pores

are relatively large, one might surmise that it might be possible to explicitly model

the flow through the porous membrane microstructure within the CFD model

itself. That is, if a geometric model of the membrane pore structures could be

developed, standard Lagrangian particle transport models included in commercial

CFD packages could be used, in conjunction with some method for modelling

particle attachment and detachment within and on the membrane microstructure,

to predict the performance of these systems. In these particle transport models,

particles are essentially injected into the CFD flow domain, and then tracked

through the domain by solving differential equations describing the particulate

position, velocity, temperature and mass [123]. However, some serious difficulties

hinder this approach:

67
x The geometry of membrane microstructures is very complex, and can vary

significantly between different membranes, as well as spatially over a

single membrane, so it is difficult to construct a geometry for general use;

x Implementing a model to describe both particle-particle and particle-

membrane interactions is very difficult; a range of models have been

developed for other process engineering applications [123] but it is

unknown how applicable these are to membrane applications;

x Finally, and most significantly, the computational expense incurred by

such simulations is prohibitively large at present. This is especially so

when considering the combination of the complexity of particle-particle

interactions, particle-membrane interactions, and the membrane

microstructure geometry.

For now, at least, this method appears mostly infeasible. Considering the higher-

pressure processes (NF and RO), on the other hand, it is more conceivable that a

general predictive model may be developed. In these processes, the feed

constituents tend to be ions or molecules dissolved into solution, and the

membrane pores are relatively small. As these pore sizes begin to approach the

scale of individual water molecules, attempting to model the microstructure of the

membrane using CFD is futile, as conventional descriptions of hydrodynamics do

not apply at these scales. Instead, a more useful approach is to use a mathematical

model to predict the rejection of the solutes within the feed solution, which is then

applied as a boundary flux condition to a CFD model of the bulk of the membrane

68
channel. In this manner, the CFD model can be used to resolve the flow structures

and concentration polarisation effects within the membrane channel, even for

complex geometries, while the rejection model can be used to describe the mass

transfer through the membrane itself. A few attempts have been made using this

coupled CFD approach to describe membrane filtration, which are summarised in

the following.

Geraldes et al. [94] used a Deen hindered transport model to predict the rejection

of various charged and uncharged compounds in single-solute solutions by a NF

membrane in conjunction with a CFD model, as previously discussed in Section

3.2.4. In this approach, the mass transfer (rejection) of the solute depends on the

local pressure differential across the membrane, obtained from the CFD solution.

However, the rejection calculation, as per Eq. (3.35), depends on the fitted

parameter (8μ/r02)(HD0ABǻP), which is constant for a given membrane-single

component solution system. This makes the approach difficult to apply generally,

and even more so when multiple-component solutions are considered.

An additional significant difficulty in modelling multiple-component solutions in

CFD is describing the diffusion of each solute throughout the solution. The

classical way to deal with this in CFD is to use a modified version of Fick’s law,

where the rate of diffusion of each component is governed by a diffusivity

coefficient, assuming that each component in the solution diffuses independently

[103, 124, 125]. However, this method does not necessarily ensure

electroneutrality of the solution in the case of charged solutes. To overcome this, a

69
Fick’s law approach can be employed assuming that salts do not dissociate, and a

mean diffusion term is used for each individual salt.

In some cases, however, the diffusion of one component may depend on other

components. For this situation, a modification of Fick’s law was put forward by

Onsager [126], using a matrix of diffusion coefficients, effectively describing the

dependence of the diffusivity of each component on all other components. While

this approach is general and theoretically sound, determination of these diffusion

coefficient matrices is very difficult, and has only been performed for a handful of

solutions.

An alternative approach is the use of the Maxwell-Stefan diffusion model [103,

125], which considers diffusion to result from the superposition of different

frictional forces between the component molecules. The formulation of the

Maxwell-Stefan approach is complex and will not be treated in detail here, but has

been used for some membrane applications with success. However, it still

effectively requires use of a matrix of diffusion coefficients which are difficult to

determine, which makes it difficult to use for a general case.

However, simplification of the Maxwell-Stefan diffusion model leads to an

approach which has already been discussed in previous sections, which is the

Nernst-Planck model (see Section 2.2.2.4). As per Fick’s law, this requires only a

single diffusion coefficient for each component, but also takes into account

electrical effects [103]. Fimbres-Weihs and Wiley [103] describe a method using

70
this approach coupled with the commercial CFD package ANSYS CFX-10.0. This

package (and most other commercial packages) assume that multicomponent

species transport occurs via Fick’s law where each component diffuses

independently. This makes it impossible to directly specify the mass fluxes for

each component directly (via the extended Nernst-Planck equations). Instead, a

mass source term must be included in the model to cancel out the Fickian

diffusion term in the default transport equations, and then add the corresponding

diffusive term from the extended Nernst-Planck equations [103]. In this manner,

the components can be allowed to diffuse based on the diffusion rates of other

components, while guaranteeing the electroneutrality of the solution is conserved.

The Fickian diffusion value used in the inbuilt transport equations is thus a

dummy value, which should not affect the simulation as the term is cancelled out

by the additional source term [103].

Fimbres-Weihs and Wiley [103] then used this method to simulate the flow of a

NaCl:KCl solution through a 2D rectangular membrane channel. No-slip

boundary conditions were enforced along the channel walls, with either a fixed

wall concentration boundary condition or a constant flux boundary condition

imposed at the membrane surface. This method yielded converged solutions

which showed different behaviour to that predicted by the classcical Fick’s Law

approach. This is reflected in Figure 3.5, which shows the difference in ionic

concentrations within the boundary layer between the solution using the extended

Nernst-Planck diffusion model and the Fickian diffusion model. However, the use

71
of the more rigorous diffusion model presented some difficulties. In particular, the

inclusion of the source term into the governing equations for the commercial code

CFX for the entire flow domain caused the mass fraction transport equations to

become numerically stiff, particularly near the membrane surface. In addition, it

was found that the value of the dummy diffusion coefficient significantly affected

the simulation results, even though the term should have simply cancelled out

[103]. It is also probable that the inclusion of the source term and the associated

expense of solving the Nernst-Planck equations would have significantly

increased the overall computational expense. For these reasons, the authors

recommended that future work in coupling a multi-component diffusion model to

a CFD simulation over the entire flow domain should avoid the use of source

terms, and focus on instead directly specifying the diffusive terms within the

inbuilt transport equations. It is noted by the present author that to the best of their

knowledge this facility is not yet available in ANSYS CFX.

72
Figure 3.5 – Difference in ionic concentrations predicted in concentration polarisation boundary layer using

both Nernst-Planck and Fickian diffusion models [103]

Déon et al. [127] also adopted a coupled model of sorts, using the DSPM-DE (i.e.

including dielectric exclusion effects) to describe the mass transfer through the

membrane, and a 2D model using the extended Nernst-Planck equations to

describe the diffusion of multiple ions through the concentration polarisation

boundary layer. This approach is similar in principle to that of [103], except that

CFD is not used to determine the hydrodynamics of the membrane channel.

Instead, classical hydrodynamic equations for turbulent pipe flow are used to

determine the velocity profiles in the concentration polarisation layer. A similar

approach was also used by Bhattacharjee et al. [128], though using the

conventional DSPM and a linear velocity profile in the concentration polarisation

73
layer. This method was able to reproduce experimental rejection data well, and

has the advantages of providing a rigorous description of the transport of ions in a

multiple-ion solution both within the membrane itself and within the

concentration polarisation boundary layer. This is an improvement over the Fick’s

law approach, however, the formulation in [127] using classical 2D velocity

profiles within the concentration polarisation layers means the model is not

completely general in terms of the hydrodynamics. That is, in complex membrane

channel geometries, the concentration polarisation boundary layer hydrodynamics

may not follow the simplifying assumptions imposed by the use of turbulent pipe

flow velocity profiles. For this reason, the more general approach including the

coupled CFD model (capable of describing the hydrodynamics for arbitrary

geometries) used by Fimbres Weihs and Wiley [103] is potentially more widely

applicable. However, this method still suffers from numerical difficulties as the

extended Nernst-Planck equations become very stiff as discussed previously.

3.5 Comments on commercial membrane simulation packages

Several commercially available software packages for simulating membrane

performance exist, generally produced ‘in-house’ by major membrane

manufacturers. The most prominent examples of this are the ROPRO® software

developed by Koch Membrane Systems, the ROSA software developed by Dow

Water & Process Solutions, the TorayDS software developed by Toray

Membrane, and the IMSDesign software developed by Hydranautics [129].

74
As these packages are generally developed ‘in-house’ and are of significant

commercial value, there is relatively little information describing the internal

computational and numerical techniques implemented in these packages.

However, these packages in general are intended to aid process engineers in

designing appropriate membrane treatment trains, i.e. large-scale configurations of

membrane modules and arrays using ‘off-the-shelf’ commercially available

membrane modules (usually those of the manufacturer supplying the software). In

general, these packages provide a design solution encompassing array sizing,

operating pressures, scaling indices, and product and concentrate water qualities,

with predictions based on nominal membrane performance under steady-state

design conditions [129]. They are not intended for more fundamental studies of

membrane performance, either in terms of hydrodynamics or membrane mass

transfer. It is emphasised that the research direction in this thesis is not aimed at

recreating a general design tool such as these commercial packages, but rather to

provide a modelling tool which may reveal some more fundamental insights into

membrane channel behaviour, which in time may be able to contribute towards

describing membrane performance under highly variable conditions in such

design tools.

3.6 Summary and directions for research

It is apparent that CFD techniques are a useful and increasingly accessible method

for describing the behaviour of membrane filtration systems. However,

75
developing a general model for the pressure-driven membrane processes (MF, UF,

NF, and RO) presents some difficulties. The differences in removal mechanisms

between the lower-pressure processes (MF and UF) and the higher-pressure

processes (NF and RO) means that a general model of pollutant removal is

perhaps unattainable at present. In particular, the particulate nature of pollutants in

MF and UF processes requires a separate and considerably more expensive

approach to CFD modelling. In addition, the fouling behaviour of such systems is

also highly difficult to predict in a general sense.

On the other hand, if non-particulate pollutants only are considered (i.e. chemical

species which are mixed within the feed solution), then the possibility of a general

model becomes greater. Such a model would in principle be capable of describing

the high-pressure processes of NF and RO, as well as some UF processes. It is

also evident that these non-particulate pollutants will be present in the low-

pressure processes of MF and UF, though their removal will be significantly

lesser and their effect on membrane performance will be small. The converse

however is not true; high-pressure processes are generally not required to treat

such large pollutants which are removed by a range of pre-treatment processes.

In addition, modelling of a wide range of membrane processes introduces the

difficult topic of turbulence. While UF and MF processes may in some instances

operate in flow regimes where turbulence is fully developed (i.e. Reynolds

numbers greater than 30 000), the majority of processes lie in lower Reynolds

number regimes, where flow is either laminar (steady), or unsteady without

76
isotropic turbulence, and hence the use of conventional turbulence models cannot

be justified. These unsteady flows can be fully resolved, however, if appropriately

fine spatial and temporal discretisation is employed, which has been shown to be

feasible for conditions encountered in NF and RO.

The use of 2D versus 3D flow models is also of interest. 3D models should be

able to more fully describe the complex flow patterns in membrane modules, and

provide the most general implementation of the model possible. However, the

great computational expense incurred by 3D models has meant that they can only

be feasibly be employed for small representative portions of the membrane

channel where spatially periodic flow conditions exist, which is not the case in

general. The use of 2D models allows computational costs to be kept to a

reasonable level such that simulations can be performed for a general case, and

also include mass-transfer processes as discussed above.

Thus, the approach adopted in this thesis has been to develop a general

computational model for the filtration of feed solutions containing dissolved

solutes only. The model will include both a detailed description of the

hydrodynamic behaviour of the membrane for laminar conditions in 2D, including

the effects of concentration polarisation, and a model of the transport of these

dissolved solutes through the membrane itself. This differs from previous studies

in which these approaches have traditionally been separate, and allows for a

reasonably complete predictive model of the high-pressure processes of NF and

RO. With selection of appropriately fine computational meshes and timestep, the

77
laminar model is also able to directly resolve unsteady flows that may occur in

these high pressure processes, without recourse to turbulence models which are

generally not appropriate for these transitional flows. The use of 2D models also

makes the combined approach computationally feasible, and not restricted to

situations where spatially periodic flow structures exist, as in most other

applications of 3D simulations.

While it is not considered in this thesis, this approach can be generalised in the

future to address particulate modelling and turbulence modelling with a view

towards application to the low-pressure processes of MF and UF. In addition, the

approach can be easily generalised in the future to include 3D modelling as the

increases in computational resources make this a feasible prospect. Finally, such a

model could be used to investigate fouling reduction techniques such as those

discussed in Section 2.3.3, though this is also not addressed in this work.

78
Chapter 4

Hydrodynamic modelling of membrane filtration

4.1 Introduction

This chapter describes the development of a basic set of rigorous hydrodynamic

models of the membrane filtration process. The models are tested and verified

against experimental and computational data from the literature, and modelling

strategies and techniques particular to the problem of membrane filtration are

discussed. Several different geometric and flow configurations are investigated,

including 2D membrane geometries with and without hydrodynamic spacer

meshes, steady-state and transient simulations, as well as steady and unsteady

flow conditions. The simulations in this chapter are concerned primarily with the

describing the hydrodynamics of the membrane channel and the phenomena of

fouling due to concentration polarisation, but do not explicitly try to simulate the

mass transfer of solutes through the membrane structure itself. Thus, the purpose

of these simulations is to establish the usefulness of CFD as a basic modelling tool

for membrane filtration, which will be extended in Chapter 6 to provide a fully

predictive model of the membrane filtration process, with the addition of a

predictive rejection model to be described in Chapter 5.

79
4.2 2D flow through a porous membrane channel

The most basic hydrodynamic model of the membrane filtration process consists

of a two-dimensional (2D) rectangular membrane channel, with one or both walls

considered to be permeable. This configuration is general enough to locally

represent a range of membrane geometries for the purpose of investigation of CP

effects. The simplification to 2D is justifiable given the low aspect ratio of most

membrane feed channels, and is considerably less computationally expensive

[130]. The software package CFX 13.0 (Ansys, Inc., USA) is used for all

simulations of the hydrodynamics of the membrane channel. CFX is a general

purpose computational fluid dynamics package, which uses an element-based

finite volume method with a co-located grid layout such that the control volumes

are identical for all transport equations. A coupled solver is used, where the

equations for continuity and momentum are solved simultaneously as a single

system [123].

CFX is, in a strict sense, capable of three-dimensional (3D) modelling only;

however, 2D models may be constructed by the use of periodic pairing, which

creates symmetry boundary conditions on the lateral faces of the model which is

then only a single element thick. This effectively creates a 2D model, and makes it

trivial to extend the model later to three dimensions as necessary.

80
4.2.1 Model geometry

The model consists of a rectangular channel as shown in Figure 4.1. The flow

through the channel is oriented along the x-axis, and enters at x = 0 through the

inlet face. The flow exits the channel at x = L through the concentrate face, and

permeation occurs through either or both of the lower channel face at y = -h and

the upper face of the channel at y = h, which represents the membrane surfaces. It

should be noted that in Figure 4.1 that permeation is only shown through the

lower channel face for clarity. The left and right ‘walls’ of the channel are coupled

under the assumption of axisymmetric flow. The right wall of the channel is

located at z = -w/2, and the left face is located at z = +w/2. The width of the

channel w (i.e. the thickness of the single width element) typically should be of

the same order as the smallest cell within the face mesh, and was varied as such

depending on the face mesh size in later grid dependence studies. While simple

rectangular prismatic elements were used for this simple geometry, the model can

be generalised to any arbitrary 2D or 3D geometry.

81
y

feed flow

z x

permeate flow

w
2h

retentate flow

Figure 4.1 - Model geometry for basic 2D model of laminar flow through porous membrane channel (not to

scale)

4.2.2 Boundary and initial conditions

Boundary conditions are specified at the channel inlet (in the form of an axial

velocity profile) and the channel outlet (as an average static pressure). It is

assumed that a fully developed laminar flow profile exists at the channel inlet.

The bulk composition of the solution is also specified at the channel inlet in terms

of a solute mass fraction for each solute i, and initial conditions for the solution

are automatically generated by the CFX solver using this specified mass fraction

over the entirety of the membrane channel.

To represent the permeate flow through the membrane, a mass sink is applied at

the membrane surface. In this case, at each computational point along the

82
membrane surface a mass sink is applied as the product of the membrane flux as

defined by Eq. (4.1) and the local density of the fluid mixture:

JV LV 'P  '3W (4.1)

where LV LV WKH K\GURG\QDPLF SHUPHDELOLW\ RI WKH PHPEUDQH DQG ǻȆW is the

osmotic pressure difference between the feed solution at the membrane surface

and the permeate. The TMP is not explicitly applied to the model, as doing so

decreases the resolution of small pressure gradients near the membrane surface

due to numerical rounding error, as has been noted previously [98]. The TMP is

instead simply applied as an input parameter to calculate the mass sink expression

representing the permeate flux. The solute mass fraction of the permeate is

specified at each computational point along the membrane surface according to

Eq. (4.2):

Zi , p Zi ,W 1  Ri (4.2)

where Ȧi,p and Ȧi,W are the mass fractions of solute i in the permeate stream, and

at the membrane surface, respectively, and Ri is the rejection of the solute i. The

rejection of a particular solute by a particular membrane is often determined

experimentally; however it would be desirable in many cases to predict the solute

rejection without recourse to experiment. A method for prediction of the

rejections Ri will be considered later in Chapter 5.

83
A symmetry condition is used for the outer ‘walls’ of the membrane channel

(those parallel to the x-y plane) to allow a 2D simulation. The remaining top and

bottom walls of the membrane channel are represented by using standard no-slip

boundary conditions in the x-direction, with a mass sink applied to represent the

permeation through the membrane surface as discussed previously. The velocity

in the y-direction is automatically determined by the CFX solver based upon the

local transport properties and the permeate flux imposed by the mass sink term,

which is a function of the local solute mass fraction and rejection.

4.2.3 Computational mesh, convergence and hydrodynamic verification

To ensure the basic hydrodynamics of the membrane channel were correct, the

model was first verified for the case where the solute has neutral properties, by

which we mean the solution viscosity, density, diffusivity, and osmotic pressure

are identical to those of water regardless of the solute concentration. The

analytical solutions of Berman [81] and Karode [85] where both upper and lower

walls of the membrane channel are permeable are used as verification cases for

the model. The model dimensions have been chosen as similar to those used by

Karode [85] for ease of verification. The relevant model input parameters are

shown in Table 4.1. The hydrodynamic behaviour of the model was verified by

observing the total axial pressure drop ǻ3axial between the membrane inlet and

outlet. The axial pressure drop was monitored to ensure that good agreement with

84
the analytical solutions was obtained, and to determine the mesh density at which

the calculated pressure drop did not change significantly.

Table 4.1 - Model input parameters for verification case

Parameter Value

Membrane channel length L 1m

Membrane channel half height h 1 mm

Membrane channel width w Variable depending on mesh density – equal to size of smallest mesh
element

Transmembrane pressure 3 bar

Membrane permeability LV 9.17 x 10-11 m Pa-1 s-1

Inlet Reynolds number NRe 250

Convergence was observed by monitoring the root mean squared (RMS) residuals

for mass, momentum, and solute mass fraction. The model was judged to be

converged once all residuals were smaller than 1 x 10-6. The hydrodynamic model

results indicated that good agreement with the analytical solutions was observed

using mesh resolutions in the order of 320-640 axial cells (in the x-direction) and

40 vertical cells (in the y-direction). The mesh size was then used as a starting

point for later refinement of the mesh needed to resolve the fine CP boundary

layer. A comparison of the non-dimensionalised axial pressure drop for the

converged numerical solution and the analytical solutions is shown in Figure 4.2.

Excellent agreement is shown between the CFD solution and the analytical

solution of Karode [85]; Berman’s solution [81] underpredicts the pressure drop

as this solution neglects the axial variation in permeate flux. The axial pressure

85
drop is, of course, lesser than that of an equivalent channel with impermeable

walls, shown for comparison.

50
CFD model
Dimensionless axial pressure drop

Impermeable walls
40
Berman (1953)
Karode (2001)
30

20

10

0
0 200 400 600 800 1000
Dimensionless axial length (x/h)

Figure 4.2 - Comparison of dimensionless axial pressure drop for CFD model with analytical solutions for

hydrodynamic verification of model

4.2.4 Incorporation of concentration polarisation effects

Concentration polarisation occurs when rejected solute accumulates next to the

membrane surface, forming a boundary layer [6]. This accumulation of solute

causes a diffusive flow back to the feed, which eventually reaches a steady-state

condition where the solute flow to the membrane surface equals the solute flux

through the membrane plus the diffusive flow from the membrane surface back to

the feed. The establishment of this boundary layer allows subsequent fouling of

86
the membrane layer to occur; minimisation or removal of the CP layer is thus a

primary goal of membrane research.

To model the effects of concentration polarisation at a basic level within the

membrane channel, the multicomponent flow option in CFX is used. This

assumes that the components of the fluid are mixed at a microscopic level so that

they share identical pressure, velocity and (where applicable) temperature fields,

and assumes that mass transfer occurs via convection and diffusion only. These

assumptions are reasonable for the majority of membrane filtration applications.

4.2.4.1 Spatial variation in transport properties

To describe the effects of concentration polarisation on the permeation of the

solute at varying points on the membrane surface, it is necessary for the transport

properties of the solution, such as diffusivity, density, and viscosity, to vary

spatially over the computational domain as a function of solute concentration A

simple approach and test case was derived from the work of Geraldes et al. [94] in

which the viscosity, diffusivity and osmotic pressure of aqueous solutions of

several compounds are described as polynomial functions of the solute mass

fraction. The compounds used in this study (sodium chloride, sodium sulphate,

sucrose and PEG1000) have been used as test compounds for the present model.

The solution density cannot be explicitly specified as a function of the solute mass

fractions due to software limitations, but is calculated using a harmonic average

87
based on the density of the individual components within the solution. The

osmotic pressure is calculated as an additional scalar parameter which varies

across the solution domain as a function of the solute mass fraction, and is then

used to determine a more accurate approximation of the permeate flux at the

membrane surface as per Eq. (4.1). The relevant expressions are shown in Table

4.2.

Table 4.2 - Trial solute components and expressions for transport properties (after [94])

Density Viscosity Diffusivity Osmotic pressure Ȇ


Compound
ȡ (g cm-³)* μ × 103 (Pa s) D × 109 (m2 sí1) × 10í5 (Pa)
1.61(1 í 14 Ȧi) for Ȧi
NaCl 2.165 0.89(1 + 1.63 Ȧi) < 0.006; 1.45 for Ȧi 805.1 Ȧi
>0.006

Na2SO4 2.664 0.89(1 + 3.52 Ȧi) 1.23(1 í 0.76 Ȧi 0.4) 337.8 Ȧi 0.95

72.18Ȧi(1 + 0.94Ȧi +
Sucrose 1.587 0.89(1 + 1.31Ȧi + 16.83Ȧ i 2) 0.52(1 í 1.33Ȧi)
2.93Ȧ i 2)

24.64 Ȧi (1 + 2.94 Ȧi
PEG1000 1.12 0.89(1 + 6.59 Ȧi + 120.8 Ȧi 3) 0.309
+ 19.25 Ȧi 2)

* Used for CFX internal calculation of solution density using harmonic average

4.2.4.2 Mesh refinement for resolution of CP layer

Significant mesh refinement is required in the vicinity of the membrane surface to

adequately resolve the fine CP boundary layer. Previous hydrodynamic models

had indicated cell aspect ratios of roughly 30 – 60 gave acceptable results. This

was used as a starting point for a grid independence study using the geometry of

Geraldes et al. [94], which is shown in Table 4.3. The vertical mesh resolution and

the degree of refinement near the membrane surface were both investigated in the

88
grid independence study. The final mesh configuration chosen was that of 100

axial cells by 75 vertical cells, with the mesh size varying in the vertical direction

such that the height of the largest mesh cell (in the centre of the channel) was 100

times that of the smallest mesh cell (adjacent to the membrane wall). An

illustrative portion of the computational mesh is shown in Figure 4.3.

(a)

(b)

Figure 4.3 - Representative detail of computational mesh for 2D channel with refinement for concentration

polarisation boundary layer at (a) channel inlet, (b) membrane surface (not to scale)

89
Table 4.3 – Model input parameters for simulations corresponding with Geraldes et al. [94]

Parameter Value

Membrane channel length L 200mm

Membrane channel half height h 1 mm

Membrane channel width w Variable depending on mesh density – equal to size of smallest mesh
element

Transmembrane pressure 1, 2, 3, 4 MPa

Membrane permeability LV 1.4 x 10-11 m Pa-1 s-1

Inlet Reynolds number NRe 500, 2000

In addition, initial model runs indicated that additional refinement was necessary

at the channel inlet to smoothly describe the ‘leading edge’ of the CP boundary

layer. The mesh was refined so that the first cells at the inlet were of unity aspect

ratio, growing progressively larger along the x-axis until x = 10 mm, after which

point the cells were of equal length for the remainder of the channel.

4.2.5 Simulations

The experimental data from Geraldes et al. [94] was used to provide the

membrane rejection factors for each solute at varying transmembrane pressures,

which are shown in Table 4.4. The membrane used in these experiments was the

CDNF50I thin-film composite NF membrane (supplied by Separem, Biela, Italy),

with a stated hydrodynamic permeability of 1.4×10-11 m2 s kg-1. Experiments were

conducted at several different crossflow velocities; the data used for this study are

for the extreme values (the lowest and highest studied velocities), which are

90
equivalent to entrance Reynolds numbers (NRe = 4u0Kȡ /Ș) of 500 and 2000,

respectively. These are equivalent to the stated Reynolds numbers of 250 and

1000 given by Geraldes et al. [94] due to a different formulation of the entrance

Reynolds number – the formulation used here is consistent with Berman [81] and

Karode [85]. The rejection values given were measured for the case where NRe =

2000.

Table 4.4 - Experimentally observed rejection factors for test compounds from [94] for NRe = 2000

7UDQVPHPEUDQHSUHVVXUHǻP
Compound 1 MPa 2 MPa 3 MPa 4 MPa
Robs* Robs Robs Robs
Sucrose 0.993 0.996 0.998 0.999

NaCl 0.403 0.572 0.649 0.715

Na2SO4 0.987 0.994 0.996 0.997

PEG1000 0.9981 0.9987 0.9993 0.9995

A comparison of the permeate fluxes predicted by the current model and the

experimental flux data and model results of Geraldes et al. [94] is shown in Figure

4.4. Values were compared at three collector sites (Collectors 1, 2 and 3) located

7.5 mm, 22.5 mm and 45 mm from the channel inlet, respectively – as no

experimental flux values were available for sodium chloride, these are not shown.

Reasonable agreement with the experimentally determined flux values is shown,

though the model consistently overpredicts the permeate flux. However, the

current model generally gives better agreement with the experimentally

91
determined fluxes than the model results of Geraldes et al. [94], except for the

case of sodium sulphate for NRe = 500.

The ability of the model to predict the formation of the CP boundary layer was

evaluated by comparing the predicted non-dimensional CP boundary layer

profiles with those presented by Geraldes et al. [94]. The non-dimensionalised CP

boundary layer height įW / h is plotted as a function of the dimensionless axial

length x / h in Figure 4.5. The profiles show good agreement with those

determined by Geraldes et al. [94]. As expected, a higher entrance Reynolds

number corresponds to a thinner CP boundary layer, due to increased shear

disrupting the CP boundary layer.

92
Sucrose, NRe = 500 Sodium sulphate, NRe = 500
6 6

Permeate flux (×10 -5 m3 m-2 s-1 ) 5 5

Permeate flux (×10 -5 m3 m-2 s-1 )


4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Transmembrane pressure (MPa) Transmembrane pressure (MPa)

PEG1000, NRe = 500 Sucrose, NRe = 2000


6 6

5 5

Permeate flux (×10 -5 m3 m-2 s-1 )


Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Sodium sulphate, NRe = 2000 PEG1000, NRe = 2000


6 6

5 5
Permeate flux (×10 -5 m3 m-2 s-1 )

Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Collector 1 model Collector 2 model Collector 3 model


Collector 1 experimental Collector 2 experimental Collector 3 experimental
Collector 1 Geraldes et al. Collector 2 Geraldes et al. Collector 3 Geraldes et al.

Figure 4.4 - Comparison of predicted permeate fluxes for CFD model with experimental and modelled fluxes

from [94] using experimentally determined rejection factors

93
Sucrose Sodium chloride
0.05 0.06

NRe = 500 NRe = 500

0.05
0.04
Dimensionless CP layer height (įW / h)

Dimensionless CP layer height (įW / h)


0.04
NRe = 2000
0.03 NRe = 2000

0.03

0.02

0.02

0.01
0.01

0 -1.39E-17
0 10 20 30 0 10 20 30

Dimensionless axial length (x/h) Dimensionless axial length (x/h)

Sodium sulphate PEG1000


0.07 0.05

NRe = 500
0.06
NRe = 500
0.04
Dimensionless CP layer height (įW / h)

Dimensionless CP layer height (įW / h)

0.05

0.03
0.04
NRe = 2000
NRe = 2000

0.03
0.02

0.02

0.01
0.01

0 0
0 10 20 30 0 10 20 30
Dimensionless axial length (x/h) Dimensionless axial length (x/h)
1 MPa model 1 MPa Geraldes et al.
2 MPa model 2 MPa Geraldes et al.
3 MPa model 3 MPa Geraldes et al.
4 MPa model 4 MPa Geraldes et al.

Figure 4.5 - Comparison of dimensionless CP boundary layer profile for CFD model with computational

results of [94] using experimentally determined rejection factors

94
Some illustrative results from the simulations are also shown here. Figure 4.6

depicts the variation in solute mass fraction over the membrane channel,

illustrating the build up of the concentration polarisation boundary layer adjacent

to the membrane surface on the bottom of the channel. The velocity field in the

membrane channel can also be seen in Figure 4.7 near the channel inlet, and in

Figure 4.8 for the entire channel. The classic laminar velocity profile is observed

within the channel, though slightly flattened towards the lower half of the channel

due to the permeation through the lower channel wall.

The variation in the solution transport properties can also be observed from the

numerical solution. Representative examples of the variation in density, viscosity,

and osmotic pressure are shown in Figure 4.9, Figure 4.10, and Figure 4.11,

respectively. The density and viscosity increase in the vicinity of the membrane

surface as expected by relatively small amounts (in the order of several percent).

The osmotic pressure variation is more pronounced however, and can vary by

approximately an order of magnitude between the bulk flow in the centre of the

channel and adjacent to the membrane surface.

95
Figure 4.6 - Representative detail of variation in solute mass fraction in 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

Figure 4.7 - Representative detail of velocity variation at inlet of 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

96
Figure 4.8 - Representative detail of velocity variation for entire 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

Figure 4.9 - Representative detail of density variation for entire 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

97
Figure 4.10 - Representative detail of viscosity variation for entire 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

Figure 4.11 - Representative detail of osmotic pressure variation for entire 2D membrane channel including

concentration polarisation boundary layer (sucrose solution, TMP 1 MPa, NRe = 500 – not to scale)

4.2.5.1 Osmotic pressure correction factor

The model consistently overpredicts the permeate flux compared to the

experimentally determined flux values. An attempt was made by Geraldes et al.

98
[94] to compensate for this over-prediction of flux by use of an osmotic pressure

correction factor ȁ, where the corrected permeate flux is given by:

J ȁ/V 'P  '3W (4.3)

To see if this approach would account for the flux discrepancy in the current

model, a set of simulations was carried out using the osmotic correction approach

with expressions for ȁ as a function of the solute mass fraction as given by

Geraldes et al. [94], which are shown in Table 4.5. No correction factors were

provided for sodium chloride.

Table 4.5 - Osmotic pressure correction factor expressions for test compounds (after [94])

Compound Osmotic pressure correction factor ȁ *

Sucrose 1 - 0.375(1000Ȧi) 0.4

Na2SO4 1 - 0.012 (1000 Ȧi) 0.785

PEG1000 1 - 0.0384 (1000Ȧi) 0.31

* Expressions have been modified to account for different units for solute mass fraction

The permeate fluxes and CP boundary layer profiles obtained from these

simulations are shown in Figure 4.12 and Figure 4.13, respectively. The overall

agreement with the experimental flux values, when averaged over all

combinations of transmembrane pressure and collector location, is improved for

PEG1000 and sodium sulphate for both inlet Reynolds numbers. However, the

overall agreement is actually decreased for sucrose for both Reynolds numbers,

99
where the model tends to underpredict the flux, especially at higher

transmembrane pressures. In all cases, however, the model’s flux predictions are

significantly closer to the experimental results than the model results of Geraldes

et al. [94]. The modelled CP boundary layer profiles remain largely identical to

those obtained without the osmotic pressure correction factor.

100
Sucrose, NRe = 500 Sodium sulphate, NRe = 500
6 6

5 5

Permeate flux (×10 -5 m3 m-2 s-1 )


Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Transmembrane pressure (MPa) Transmembrane pressure (MPa)

PEG1000, NRe = 500 Sucrose, NRe = 2000


6 6

5 5

Permeate flux (×10 -5 m3 m-2 s-1 )


Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Sodium sulphate, NRe = 2000 PEG1000, NRe = 2000


6 6

5 5
Permeate flux (×10 -5 m3 m-2 s-1 )
Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Collector 1 model Collector 2 model Collector 3 model


Collector 1 experimental Collector 2 experimental Collector 3 experimental
Collector 1 Geraldes et al. Collector 2 Geraldes et al. Collector 3 Geraldes et al.

Figure 4.12 - Comparison of predicted permeate fluxes for CFD model incorporating osmotic pressure

correction factor with experimental and modelled fluxes from [94] using experimentally determined rejection

factors

101
Sucrose Sodium chloride

0.05 0.06

NRe = 500 NRe = 500

0.05
0.04
Dimensionless CP layer height (įW / h)

Dimensionless CP layer height (įW / h)


0.04
NRe = 2000
0.03 NRe = 2000

0.03

0.02

0.02

0.01
0.01

-2.78E-17 0
0 10 20 30 0 10 20 30
Dimensionless axial length (x/h) Dimensionless axial length (x/h)

Sodium sulphate PEG1000


0.07 0.05

NRe = 500
0.06
NRe = 500
0.04
Dimensionless CP layer height (įW / h)

Dimensionless CP layer height (įW / h)

0.05

0.03
0.04
NRe = 2000
NRe = 2000

0.03
0.02

0.02

0.01

0.01

0 -2.78E-17
0 10 20 30 0 10 20 30

Dimensionless axial length (x/h) Dimensionless axial length (x/h)

1 MPa model 1 MPa Geraldes et al.


2 MPa model 2 MPa Geraldes et al.
3 MPa model 3 MPa Geraldes et al.
4 MPa model 4 MPa Geraldes et al.

Figure 4.13 - Comparison of dimensionless CP boundary layer profile for CFD model incorporating osmotic

pressure correction factor with computational results of [94] using experimentally determined rejection

factors

102
4.2.6 Concluding remarks

It can be seen from comparison to analytical, experimental and computational data

that the present CFD modelling approach is capable of describing the flow

conditions for 2D membrane channels. Good agreement is observed in terms of

axial pressure drop, permeate flux, and the formation of the concentration

polarisation boundary layer. This gives reassurance that the basic hydrodynamic

model of the membrane filtration process is satisfactory, and we may now turn our

attention to more complex geometrical and hydrodynamic configurations. We first

consider the case of flow in spacer-filled channels.

4.3 Flow in spacer-filled channels

One area in which CFD methods have been reasonably prominent is in the

modelling of spiral-wound membrane (SWM) modules, as evidenced by a large

body of recent work in the area [36, 47, 113, 115, 119, 122, 131-136]. These

modules can exhibit complex hydrodynamic and mass transfer phenomena,

largely due to the effects of spacer meshes within the modules, which in addition

to separating the leaves of the membrane, are able to reduce the occurrence of CP

by promoting mixing and flow instabilities at the membrane surface, and hence

increase mass transfer through the membrane. This increase in mass transfer

comes at the expense of energy loss, though frequently the economic benefit of

increased mass transfer is greater than the disadvantage of energy loss. As direct

experimental observation of the flow field within SWM modules is difficult

103
without disturbing the flow structures of interest, CFD has become a powerful

tool for investigating the trade-off between optimisation of mass transfer and

minimisation of energy loss. Some of the primary considerations for CFD

modelling of spacer-filled channels are briefly reviewed in the following.

4.3.1 Unsteady flow and turbulence in spacer-filled channels

As discussed in Section 3.3, unsteady flow patterns can be observed in many

membrane filtration applications. This is particularly the case in spacer-filled

channels, especially at higher Reynolds numbers, as has been observed by several

authors [42, 137]. A casual evaluation of this might lead to the conclusion that use

of a turbulent CFD model may be beneficial for these cases. However, true fully-

developed, isotropic turbulence (generally for Reynolds numbers greater than 30

000) is only likely to be observed in the lower pressure membrane filtration

processes of MF and UF, and not for the high pressure NF and RO processes. In

particular, Reynolds numbers for SWM modules tend to be in the range 1000-

3000 [42, 100, 103]. Because of this, turbulence models (such as the common k-İ

model) which rely on the assumption of fully developed isotropic turbulence are

unsuitable for these conditions. Instead, a common approach is a form of direct

numerical simulation, using laminar flow models for transient simulation using

very fine computational meshes and timesteps [e.g. 42, 103, 111-114, 115, 137].

This is the approach that has been adopted in the subsequent simulations

discussed in Section 4.3.7.

104
4.3.2 2D and 3D modelling

As discussed in Section 3.3.2, both 2D and 3D models have been employed with

success to describe unsteady flow in spacer-filled membrane channels. However,

due to computational expense, 3D models are generally only currently feasible

using a unit-cell approach with ‘wrapped’ spatially periodic boundary conditions,

and with a simplified dissolving-wall boundary condition to represent membrane

mass transfer. The addition of a true permeable wall boundary condition adds

considerable computational expense, and is rare in the literature for 3D modelling.

With this in mind, a 2D model has been adopted for this investigation to reduce

computational requirements to manageable levels, and allow for the incorporation

of the permeable wall condition as employed in previous simulations. This also

allows for the further incorporation of predictive mass transfer modelling in a

coupled simulation approach (to be discussed in Chapter 6) within reasonable

computational constraints.

4.3.3 Model geometry

The model consists of a rectangular 2D channel filled with a ‘zigzag’ cylindrical

spacer arrangement identical to that used by Wardeh and Morvan [115], shown in

Figure 4.14. This zigzag arrangement is a reasonable 2D representation of a

woven spacer net, which is one of the most common spacer geometries. The flow

through the channel is oriented along the x-axis, and travels through an entrance

length equal to 12 times the spacer filament diameter df. It then travels past seven

105
spacer filaments which are followed by an exit length equal to 25 times the spacer

filament diameter. The filaments are separated by an axial distance of l, and the

channel height is h. For all simulations, the following configuration was used: h =

2 mm, l/h = 4, and df / h = 0.5.

The large entrance and exit lengths are necessary so that flow is fully developed

before the spacer-containing unit cells, and so that the outlet condition does not

impinge on the recirculation zones created by the final spacer [114]. The use of

only seven spacer filaments is a computationally economic way to represent the

spatially periodic velocity and pressure field in the spacer channel which in

practice is much longer. Simulations by Fimbres-Weihs et al. [114] showed at

least six filaments were necessary to obtain vortices in unsteady flow, and that

flow around the last spacer filament exhibited different behaviour due to exit

effects. In light of this, the region in the vicinity of the fifth and sixth spacers is

used as a representative monitoring location, as it is far enough downstream for

the periodic velocity and pressure flow structures to become established, but

sufficiently removed from exit effects as well [114].

spacer filaments
channel height h
diameter df permeate flow

feed retentate

entrance length permeate flow exit length


12 df 25 df

inter-filament distance l

Figure 4.14 – Model geometry for simulations of flow through 2D space-filled channel

106
4.3.4 Boundary and initial conditions

Boundary conditions are specified at the channel inlet (in the form of a uniform

axial velocity) and the channel outlet (as an average static pressure of zero). The

average inlet velocity uavg is specified in terms of a channel Reynolds number as

used by several authors [44, 111, 114], which is given by:

Uuavg d h
Re ch (4.4)
K

where ȡ is the fluid density, Ș is the fluid viscosity, and dh is the hydraulic

diameter for a spacer-filled flow channel:

4H
dh
2 hch  1  H 4 d f
(4.5)

and where İ is the voidage of the spacer-filled channel:

Vtot  VSP
H (4.6)
Vtot

where Vtot and VSP are the total volume per unit cell and volume occupied by

spacer filaments per unit cell, respectively. The model was run for four different

Reynolds numbers of 100, 200, 400 and 800 as per the work of Wardeh and

Morvan [115].

No-slip boundary conditions with no mass transfer are used along the channel

walls over the entrance and exit lengths, as well as on the curved surfaces of the

107
membrane spacers. The bulk composition of the solution is specified at the

channel inlet in terms of a solute mass fraction for each solute i; here, a simple

salt solution (sodium chloride) is considered. To represent the permeate flow

through the remainder of the membrane surface, a mass sink is again applied at

the membrane surface as previously described in Section 4.2.2. A constant

rejection value for the salt is used for all simulations. The relevant parameters for

the simulations are shown in Table 4.6, as per [115].

Table 4.6 – Model input parameters for simulations corresponding with Wardeh & Morvan [115]

Parameter Value

Transmembrane pressure 0.5 MPa

Membrane permeability LV 2 x 10-10 m Pa-1 s-1

Channel Reynolds number Rech 100, 200, 400, 800

Salt rejection R 0.95

Initial conditions are again automatically generated by the CFX solver, with the

initial solute concentration field set equal to that of the bulk solute mass fraction

throughout the domain.

4.3.5 Incorporation of concentration polarisation effects

As per Section 4.2.4.1, 87 the effects of concentration polarisation were included

by allowing the transport properties of the solution, such as diffusivity, density,

and viscosity, to vary spatially over the computational domain as a function of

solute concentration, using the relations provided by Geraldes et al. [94]. The

108
same approach was adopted by Wardeh and Morvan [115], though they were able

to directly specify the density using the relationship provided in [94]. This was not

possible in the present simulations, and again the harmonic average approach to

determine the solution density was employed as described in Section 4.2.4.1.

4.3.6 Computational mesh and convergence criteria

Due to the more complex geometry of the spacer-filled channel, unstructured

meshes were used for all simulations of the spacer-filled channel. As before,

significant mesh refinement is necessary in the vicinity of the membrane surface

to resolve the concentration polarisation boundary layer, and in this case

additional refinement is also necessary adjacent to the spacer filament surfaces. A

representative detail of a typical computational mesh in the vicinity of a spacer

filament is shown in Figure 4.15.

A mesh dependence study was conducted using a series of progressively finer

meshes to determine the mesh density after which variation in the solution was

considered to be negligible. The adopted mesh uses an element size of

approximately 0.05 mm in the bulk of the channel, growing progressively smaller

in the vicinity of the membrane and filament surfaces, where the smallest element

size is approximately 3 μm. This is similar to the mesh used by Wardeh and

Morvan [115], with a minimum element size of 5 μm and an element size in the

order of 0.02 mm in the bulk of the channel.

109
Convergence was again observed by monitoring the root mean squared (RMS)

residuals for mass, momentum, and solute mass fraction. The model was judged

to be converged once all residuals were smaller than 1 x 10-6. It should be noted

this is more stringent than the convergence criteria chosen in [115] (a RMS

residual value of 1 × 10-5).

Figure 4.15 – Representative detail of computational mesh for 2D spacer-filled channel CFD model

110
4.3.7 Transient simulations

A set of transient simulations were run for the conditions described in Table 4.6 to

verify the model against the results of Wardeh and Morvan [115]. Previously

steady-state simulations only were conducted in Section 4.2, as the relatively low

Reynolds numbers and simple channel geometry therein led to steady (invariant

with time) solutions being readily attainable. However, with the more complex

geometry of the spacer-filled channel and the comparatively large Reynolds

numbers, unsteady flow structures (i.e. varying with time) were found to exist by

Wardeh and Morvan [115], and also for similar conditions by other researchers

[e.g. 42, 113, 114]. For this reason, transient simulations were performed with

small time steps to accurately capture the unsteady flow behaviour. As per [114],

an approximate steady-state solution was chosen as the initial state for the

transient simulations.

Very small time steps are required to adequately resolve the unsteady flow

behaviour in such channels. A common way of representing this requirement is by

considering a dimensionless time step tstuavg / df; Koutsou et al. [42] used

dimensionless time steps smaller than 1.5 × 10-3, while Fimbres-Weihs et al. [114]

used time steps as small as 8.6 × 10-4. In light of this, the dimensionless time step

adopted for these simulations was set to the smallest dimensionless time step

found in the literature (8.6 × 10-4), corresponding to a physical time step varying

from 3.17 × 10-4 s with Rech = 100 to 3.96 × 10-5 s with Rech = 800. This is smaller

than the timestep chosen by Wardeh and Morvan [115] of 1 × 10-4 s with Rech =

111
800, but is reasonably similar. The simulations were run for slightly greater than

two residence times as per [114]. Because of the fine mesh and very small

timestep, these simulations incur considerable computational time (in the order of

weeks on an eight-CPU 64 bit machine with 16 GB of RAM).

Another way of evaluating the timestep selection is by examination of the

simulation Courant number. In all cases the chosen time step yielded a RMS

Courant number of approximately 0.2, with the maximum Courant number

observed less than 2.4. This is consistent with the values obtained in [114].

The results were verified against computational results presented in [115] for the

axial pressure drop, area-averaged wall shear stress, and area-averaged salt mass

fraction on the membrane surface in the spacer-filled channel. These results are

shown in Figure 4.16, Figure 4.17, and Figure 4.18, respectively.

Generally, good agreement is observed between the current simulation results and

that of Wardeh and Morvan [115]. The predicted variation in axial pressure drop

(shown in Figure 4.16) and wall shear stress (shown in Figure 4.17) in particular

is very close to that predicted in [115], though the current model predicts slightly

greater values of both. However, when considering the area-averaged salt mass

fractions (shown in Figure 4.18), the values predicted by the current model are

lesser, particularly at higher channel Reynolds numbers. The discrepancy between

the model predictions is thought be to due to several factors, including:

112
x The different treatment in specifying the variable density of the salt

solution, as discussed in Section 4.3.5;

x The small differences in computational mesh and timestep selection (i.e. a

smaller timestep was used for the present model, and smaller mesh sizes

adjacent to the membrane wall were used for the present model); and

x The differences in convergence criteria used (i.e. the current model has a

more stringent set of convergence criteria).

350

300
Axial pressure drop (Pa)

Current model
250 Wardeh
200

150

100

50

0
0 100 200 300 400 500 600 700 800 900

Channel Reynolds number

Figure 4.16 – Comparison of predicted axial pressure drop as a function of channel Reynolds number against

computational results of [115] for spacer-filled channel

113
1.4

1.2 Current model


Wall shear stress (Pa)

1 Wardeh
0.8

0.6

0.4

0.2

0
0 50 100 150 200 250 300 350

Axial pressure drop (Pa)

Figure 4.17 – Comparison of predicted area-averaged wall shear stress as a function of axial pressure drop

against computational results of [115] for spacer-filled channel

0.006

0.005
Salt mass fraction

0.004

0.003
Current model
0.002 Wardeh

0.001

0
0 100 200 300 400 500 600 700 800 900

Channel Reynolds number

Figure 4.18 – Comparison of predicted area-averaged salt mass fraction as a function of channel Reynolds

number against computational results of [115] for spacer-filled channel

In addition, the comparative decrease in mass fraction adjacent to the membrane

surface is qualitatively commensurate with the comparative increase in the wall

shear stress compared to the computational results in [115]. Considering these

factors, it may reasonably be judged that the model is producing reasonable

results and can be considered to be verified.

114
4.3.7.1 Steady flow structures

It is interesting to now examine the transient model results to observe the

evolution of any unsteady flow structures (if they exist) over time. Beginning first

with the lowest channel Reynolds number (Rech = 100), steady flow was observed

throughout the simulation period. That is, as time progressed through the

simulation period, there was no observable variation in the calculated flow

variable fields. The same was also true for channel Reynolds numbers of 200 and

400. This is consistent with previous computational results in the literature for

similar conditions [e.g. 111, 114].

Some representative graphs of the steady-state solutions for these lower channel

Reynolds number flows are shown in Figure 4.19 through Figure 4.22, all in the

vicinity of the fifth and sixth spacer filaments (chosen as a representative

monitoring location as discussed previously). In all instances (and all subsequent

graphical representations of the 2D spacer-filled channel), the bulk flow is from

left to right. Figure 4.19 shows the variation in salt mass fraction throughout the

channel, and clearly demonstrates the disruption of the concentration polarisation

layer due to the increased shear and recirculation induced by the spacer filaments.

The flow structures are the same as that described previously in the literature for

‘zig-zag’ channels [111, 112, 114], where the flow follows the zig-zag pattern

imposed by the spacers. The influence of increasing crossflow velocity (expressed

in terms of increasing channel Reynolds number) can also be seen in Figure

4.19;the formation of high concentration areas behind the filaments is very

115
apparent at Rech = 100. However, as the channel Reynolds number increases, the

increased recirculation and shear reduces the formation of these high

concentration zones. However, at higher channel Reynolds numbers, similar zones

are formed at the leading edge of the filaments due to recirculation.

Rech = 100

Rech = 200

Rech = 400

Figure 4.19 – Predicted salt mass fraction distribution for spacer-filled channel for conditions in [115] for

steady flow at lower channel Reynolds numbers

This can also be seen via examination of the fluid streamlines and velocity vector

plots, shown in Figure 4.20 and Figure 4.21 respectively. Large recirculation

zones are apparent on the downstream faces of the filaments, while smaller

116
recirculation zones occur on the upstream faces adjacent to the membrane walls.

At Rech = 400, the downstream recirculation zone extends nearly to the next

filament, growing progressively smaller with smaller channel Reynolds number.

Rech = 100

Rech = 200

Rech = 400

Figure 4.20 – Predicted fluid streamlines for spacer-filled channel for conditions in [115] for steady flow at

lower channel Reynolds numbers

117
Rech = 100

Rech = 200

Rech = 400

Figure 4.21 – Predicted velocity vectors for spacer-filled channel for conditions in [115] for steady flow at

lower channel Reynolds numbers

This can also be visualised in terms of the vorticity, which is essentially the

circulation per unit area at each point in the flow domain. A plot of the vorticity in

the monitoring region (the fifth and sixth filaments) is shown in Figure 4.22. The

figure shows areas of high vorticity on the leading edge of the filaments in the

vicinity of the channel centre, which become increasingly pronounced with

increased channel Reynolds number. The recirculation areas behind filaments are

also clearly visible.

118
Rech = 100

Rech = 200

Rech = 400

Figure 4.22 – Predicted vorticity fields for spacer-filled channel for conditions in [115] for steady flow at

lower channel Reynolds numbers

4.3.7.2 Unsteady flow structures

For the highest channel Reynolds number studied (Rech = 800), significant

unsteady flow behaviour was observed (i.e. the computed mass fraction, velocity,

and pressure fields varied throughout the simulation time). The existence of

unsteady flow at this channel Reynolds number is consistent with previous CFD

results in the literature [e.g. 114].

119
We can first examine the evolution of the salt mass fraction distribution over time

within the membrane channel, shown in Figure 4.23. The computed mass fraction

fields are shown at time increments of approximately 6.1 ms from top to bottom

(similar to that used in [114] to visualise such flows), and partially illustrate the

movement of vortices through the spacer-filled channel in this unsteady flow. A

fuller understanding of the flow can be obtained by examining these in

conjunction with successive plots of fluid streamlines (shown in Figure 4.24),

velocity vectors (shown in Figure 4.25), and vorticity (shown in Figure 4.26), at

identical time intervals.

In a manner similar to [114], it is useful to identify and numerically label some of

the individual vortices to track their movement through the spacer-filled channel

over time. Specifically, individual vortices are assigned a number when they

appear at the leftmost portion of the monitoring region, and this number is used to

track the movement of each particular vortex as it moves downstream (from left to

right) through the monitoring region with time (from the top to bottom of the

page). (It is not intended to perform an exhaustive study of vortex shedding

behaviour in spacer-filled channels, as excellent examples of this can be found in

other papers [e.g. 111, 112, 114, 137]. Instead, we merely provide a qualitative

demonstration of the phenomenon to show some salient features of spacer-filled

channel flow).

The movement of vortices is perhaps most easily visualised via examination of the

streamlines in Figure 4.24. Vortices are formed both in front of filaments (e.g.

120
vortices 4 and 6), and behind filaments adjacent to the membrane wall (e.g. vortex

7). Vortices formed at the leading edge of filaments are eventually ‘squeezed’

over the filament in a reasonably periodic fashion, where they form elongated

elliptical vortices adjacent to the main streamflow (e.g. vortices 1, 2, and 3).

These then travel downstream between filaments and coalesce (seen with vortices

2 and 3) before eventually reaching the next filament and ‘squeezing’ past.

Vortices formed behind filaments adjacent to the membrane wall (vortex 7, for

example) appear to also travel downstream below the larger vortices adjacent to

the main streamflow, and coalesce approximately halfway between filaments

adjacent to the membrane wall.

Examination of the velocity vector plots in Figure 4.25 also reveals that vortices

also rotate in different directions, depending on whether they are above or below

the main streamflow. Specifically, those vortices on the lower side of the channel

rotate clockwise, and those on the upper side rotate counter-clockwise. This is

consistent with the behaviour observed in [114] for a similar channel, and results

in the wall shear generated by these vortices being of opposite sign to the shear

generated by the main channel flow, as noted in [114].

The shearing behaviour of these flows has some effects on the concentration

polarisation layer and hence the salt distribution, as seen in Figure 4.23. The

greatest disruption to the concentration polarisation layer occurs directly opposite

the filaments on the opposing wall of the membrane, corresponding with the areas

of highest vorticity as shown in Figure 4.26. Regions of high salt concentration

121
also build up behind filaments as observed for the steady flow cases at lower

channel Reynolds number. However, the high concentration regions in front of the

filaments observed for low Rech in Figure 4.19 are not as prevalent for the

unsteady flow case, due to the vortex-induced movement of fluid, also consistent

with the observations in [114].

122
q
n o
p

r
q
n
s o,p

r
q
s n
o,p

r
s
n
t

s
n
t

Figure 4.23 – Predicted salt mass fraction distribution for spacer-filled channel for conditions in [115] for

unsteady flow at high channel Reynolds number (Rech = 800) at successive 6.1 ms timesteps (top to bottom)

123
q
n o
p

r
q
n
s o,p

r
q
s n
o,p

r
s
n
t

s
n
t

Figure 4.24 – Predicted fluid streamlines for spacer-filled channel for conditions in [115] for unsteady flow at

high channel Reynolds number (Rech = 800) at successive 6.1 ms timesteps (top to bottom)

124
q
n o
p

r
q
n
s o,p

r
q
s n
o,p

r
s
n
t

s
n
t

Figure 4.25 – Predicted velocity vectors for spacer-filled channel for conditions in [115] for unsteady flow at

high channel Reynolds number (Rech = 800) at successive 6.1 ms timesteps (top to bottom)

125
q
n o
p

n q
o,p
s

r
q
s n
o,p

r
s
n
t

s
n
t

Figure 4.26 – Predicted vorticity fields for spacer-filled channel for conditions in [115] for unsteady flow at

high channel Reynolds number (Rech = 800) at successive 6.1 ms timesteps (top to bottom)

126
4.3.8 Concluding remarks

The results presented for 2D flow in spacer-filled channels indicate that the

current model agrees well with computational data presented in the literature. In

addition, the current model includes a more physically realistic representation of

the flow by incorporating a true permeable wall boundary condition for the

membrane surfaces, which has largely been neglected by other researchers for

spacer-filled channel flow. The model results have also demonstrated the ability

of the transient laminar flow model to describe unsteady flows for moderate

channel Reynolds numbers, when used with appropriately fine spatial and

temporal discretisation. Using such a model, unsteady flow features are observed

at channel Reynolds of Rech = 800, and formation and movement of vortices can

be clearly observed in the membrane channel, consistent with previous results

presented in the literature. However, significant computational expense is incurred

by directly simulating the unsteady flow structures using the laminar flow model,

with model run times in the order of weeks on an eight-CPU 64 bit machine with

16 GB of RAM.

4.4 Additional remarks on 3D modelling

Previous sections have touched upon the large computational expense required for

3D modelling as opposed to 2D modelling. Given the not insignificant expense of

the transient simulation of unsteady 2D flows discussed in Section 4.3, it is

perhaps useful to comment further on the feasibility of 3D simulations.

127
The high cost of 3D simulations (generally only used for simulation of spacer-

filled channels) is largely compensated for by use of a unit-cell approach with

spatially periodic boundary conditions, as discussed previously. The assumption

of spatially periodic flow that this method relies on is valid in many spacer-filled

channels, though not in some instances. As part of this study, some preliminary

investigation into the computational and mesh requirements for 3D modelling was

undertaken, which is briefly presented here to give some context to, and

justification of the adoption of 2D models for this thesis.

The membrane channel configuration modelled by Wardeh and Morvan in 2D

[115], which has been reproduced and modelled in detail in Section 4.3, provides

a reasonable comparison point for the use of 2D and 3D models. To investigate

computational requirements for extending this simulation to a 3D case, model

geometries and meshes were constructed for the same channel geometry described

in Section 4.3.3, by assuming a 3D woven spacer net geometry. This was

conducted for two scenarios:

x a unit-cell approach, where the 3D computational domain extended for

two inter-filament distances in the x and z directions; and

x an entire channel approach, where the 3D computational domain was

extended for the entire channel length (i.e. over seven spacer filaments) as

modelled in the 2D case (including entrance and exit lengths), in the x

direction, and for two inter-filament distances in the z direction.

128
Unstructured meshes were generated for both of these geometries to examine

potential memory requirements and feasibility of performing 3D simulations for

these geometries. Meshes were generated on a dedicated 64-bit machine (8

processors, 16 GB of RAM), with considerably more computational power than a

typical desktop computer. It is useful to consider as a reference point the mesh

sizes established by the 2D grid independence study (mesh elements of 0.05 mm

in the bulk of the channel, progressively refined to elements of 3 μm in areas of

interest). To minimise memory requirements, meshes can be generated only for

the quarter unit cell geometry depicted in Figure 4.27, and the symmetry of the

unit cell thus exploited to construct the entire geometries by ‘stitching’ mesh

sections together. Even with this technique, an equivalently fine spatial resolution

for the unstructured 3D mesh was not achievable on this machine, as the meshing

process exceeded the memory limits of the machine for even the small unit cell

geometry.

The finest mesh that could be generated contained mesh elements of 0.05 mm in

the bulk flow with elements of 7.5 μm adjacent to the membrane surfaces. This

represents a mesh consisting of approximately 65.7 million elements, which by

current standards is a very large mesh. For comparison, the 2D mesh adopted in

Section 4.3.6 consists of approximately 295 000 elements; and 3D meshes for unit

cell problems found in the literature range from approximately this size (e.g.

[122]) up to large meshes of 15-20 million elements [118].

129
Figure 4.27 – Quarter unit-cell geometry corresponding to 2D model dimensions used in [115]

A coarser mesh was thus constructed with mesh elements of 0.1 mm in the bulk,

and wall elements of 20 μm, to yield a total mesh size of 9.7 million elements.

This is in the general range of mesh sizes presently investigated in the membrane

literature. This coarse mesh is shown in full in Figure 4.28, with a representative

detail shown in Figure 4.29.

130
Figure 4.28 – Coarse 3D mesh (9.7 million elements) for unit-cell corresponding to 2D model dimensions

used in [115]

Figure 4.29 – Representative detail of coarse 3D mesh (9.7 million elements) for unit-cell corresponding to

2D model dimensions used in [115]

131
A test case was run for the unit-cell model using the conditions previously

modelled in Section 4.3 with the coarse 3D mesh described above. Periodic

boundary conditions were established at the boundaries of the unit cell as

described in the literature by various authors [118, 121, 122]. As mentioned

previously, the majority of 3D unit cell models have considered the simplified

cases of impermeable membrane walls or the dissolving wall approximation. As a

first attempt, simulations were run using an impermeable wall boundary

conditions for a channel Reynolds number of 100. Using this approach, a well-

converged steady-state solution was able to be obtained. However, no detailed

results are presented, as the results could not be verified using a grid dependence

study due to the constraints on computational time.

However, applying the permeable wall boundary condition used in 2D simulations

led to convergence difficulties for the mass fraction equations, and a reliable

converged solution was not able to be obtained. It is probable that a more refined

mesh could in this case lead to a converged solution, but again this is limited by

the extensive computational cost, and no further efforts were expended in this area

due to time considerations.

Perhaps most importantly, computational times for the 3D steady-state solution

with impermeable walls were in the order of hundreds of times larger than for an

equivalent 2D steady-state solution. When one considers that the computational

times required for transient simulations for the 2D channel considered in Section

4.3.7 with permeable walls were in the order of weeks, it becomes apparent that

132
3D simulations with permeable walls would require very large computational

resources to be feasible.

Similarly, the larger model geometry required for the entire channel approach

proved prohibitively large for simulation at the same mesh resolution. It was not

possible to run these simulations on the computational facilities available due to

memory requirements, regardless of the boundary conditions adopted (permeable

or impermeable walls).

Hence, it may be concluded that for most practical purposes, full 3D simulations

of a membrane channel are as yet still not possible, and the 3D unit cell approach

remains the only viable option for the near future. However, this still imposes

limits on the level of mesh refinement available; while converged solutions may

be obtained using the simplifying assumption of an impermeable wall boundary

condition, convergence using a permeable wall boundary condition has proved to

be difficult. It appears that very fine mesh resolutions beyond the present

computational facilities available for this project are required to obtain converged

solutions when considering a true permeable wall boundary condition. Further, the

computational time requirements for such an undertaking are also prohibitively

large for most purposes, and indeed for the purposes of this thesis, at present.

133
4.5 Conclusions

The use of multicomponent CFD models using the commercial CFD code Ansys

CFX has been shown to well describe the hydrodynamic behaviour of membrane

channels. The inclusion of a true permeable wall boundary condition in these

models, rather than the impermeable wall or dissolving-wall approximations most

commonly encountered in the literature, has also been shown to match well with

experimental, analytical, and numerical predictions. This also allows for the

permeation rate to vary spatially throughout the membrane channel, introducing

an additional level of detail into the hydrodynamic predictions.

Concentration polarisation effects can also be well described using such

multicomponent models by allowing the transport properties of the solution

(viscosity, density, diffusivity, and osmotic pressure) to vary spatially according

to the local concentrations of each solution component. This is achieved by using

empirical relationships derived from the literature for the certain components

considered in this study; however, it is noted that a more general method of

predicting these properties would undoubtedly be beneficial. Comparison to

predicted concentration polarisation layer thicknesses from the literature produced

favourable results, and the model was hence able to predict the corresponding

(experimentally measured) reduction in permeate flux reasonably accurately.

Unsteady (time-varying) flow for the moderate channel Reynolds numbers

observed in high-pressure membrane filtration can also be well described by use

134
of transient laminar flow models, with appropriately fine temporal and spatial

discretisation to directly resolve unsteady flow structures (in effect, a form of

direct numerical simulation). This method was shown to agree well with

computational results previously presented in the literature, and produced

qualitatively similar behaviour to previous studies in terms of vortex formation

and transport for higher channel Reynolds numbers. Importantly, the transient

laminar flow approach is well suited to describing high-pressure membrane

filtration processes, where Reynolds numbers fall well below the generally

DFFHSWHGOLPLWIRUWKHXVHRIWXUEXOHQFHPRGHOV 5H§ 

Finally, the models described in this chapter are suitable for implementation in

either 2D or 3D geometries, though simulations have been performed only for 2D.

The primary reason for this is computational expense, especially when the direct

simulation of unsteady flow effects using a transient laminar flow scheme is

considered. However, extension of the modelling approach to 3D is relatively

straightforward, but limited by present computational capabilities.

In summary, the CFD methods described here can well represent the

hydrodynamics of a membrane channel. However, they do not provide any

capacity for predicting the selective transfer of mass through the membrane itself

(i.e. the rejection of solutes), arguably the most important characteristic of a

membrane filtration system. Coupling of these hydrodynamic models to models of

the membrane mass transfer process would therefore provide a more complete

135
predictive model of the overall process. We now consider the numerical

implementation of a mass transfer model for this precise purpose in Chapter 5.

136
Chapter 5

Prediction of membrane rejection

5.1 Introduction

The primary purpose of membrane filtration is the separation of components from

a feed stream, which we have previously characterised using the concept of

rejection. The rejection of various components however can vary significantly,

depending on the nature of the components themselves, the nature of the

membrane, and the operating conditions of the filtration system, such as the

crossflow velocity, applied pressure etc. This great variability has meant that

historically, the most practical method for determining the rejection of a

component or components in a membrane filtration system has been through

experimental measurement, which may be expensive or impractical.

It is apparent that for a computational model of the membrane filtration process to

have general utility, a general method for predicting the rejection of feed streams

containing an arbitrary mixture of components, with an arbitrary membrane, and

arbitrary operating conditions is needed. The following chapter describes the

application of a computational method for this purpose, largely with regard to NF

and RO processes. This computational approach is then incorporated into rigorous

hydrodynamic simulations of the membrane filtration process, to be discussed in

detail in Chapter 6.

137
5.2 Approaches to describe separation in NF and RO processes

Two main descriptions of the solute separation process in NF and RO have

historically been used: the first are based on phenomenological (non-structural)

models in which the membrane is treated as a ‘black box’, and no characterisation

of the structural and electrochemical properties is required. Examples of this

approach include the early work of Spiegler and Kedem [138] and Kedem and

Katchalsky [139], and this approach has still been used in recent times [140]. In

these methods, membrane performance is characterised by use of fitting

parameters (salt permeability and a reflection coefficient) [e.g. 138, 139, 140].

These methods are limited in that they are only able to describe the behaviour of

binary salt systems (or in the limiting case, binary salt systems in the presence of a

completely rejected uncharged solute) [141].

The second approach are commonly termed ‘pore models’, or structural models,

and describe separation in terms of hydrodynamic permeation through a set of

(usually cylindrical) pores, in which the convective and diffusive transport of

solute is hindered by interactions between the solute and the membrane. It should

be noted that the use of pore models for NF and RO membranes does not imply

the physical existence of geometrically well-defined pores in these membranes;

rather, membranes can be characterised by an effective pore size which

approximates the hindrance in solute transport caused by the complex polymeric

microstructure of these membranes [17]. These models were initially proposed to

138
describe rejection of uncharged solutes in NF processes, but can be generalised to

describe transport of charged and uncharged solutes in most membrane types.

This is usually done by solving the extended Nernst-Planck equations using the

Donnan steric partitioning model (DSPM) to describe the transport of solutes

through the membrane [e.g. 16, 17, 127, 142, 143-150]. This approach has several

advantages: first, it is based on a more rigorous description of physical parameters

of the membrane system only, and all parameters used have a reasonably direct

physical interpretation. Secondly, it is capable of describing the transport of both

charged and uncharged solutes, and mixtures thereof.

However, the DSPM has been noted to have several limitations. In particular, it

has been shown that in some instances that the method fails to accurately predict

the rejection for divalent counterions [151]. Additionally, very high values for the

membrane charge can be obtained when used as a fitting parameter, which may

not be physically reasonable [e.g. 152, 153]. For this reason extensions to the

DSPM using the concept of dielectric exclusion have been proposed by several

authors [18, 19, 151], generally referred to as the DSPM-DE model. The dielectric

exclusion effect is caused by two factors: first, the difference in dielectric constant

between an aqueous solution and the surrounding membrane structure [151]; and

additionally, the difference in dielectric constant of the solution between the bulk

of the solution and in the confined membrane pores [154]. The DSPM-DE model

has indeed been shown to better predict rejection of multivalent counterions [e.g.

19, 151].

139
Despite its limitations, however, the DSPM represents a reasonable starting point

for developing a general numerical method to predict rejection, which will be

applied later (see Chapter 6) in conjunction with a CFD model of the membrane

channel hydrodynamics. For this reason, the extensions to the DSPM provided by

the dielectric exclusion models have been neglected for the current study, but

represent a useful and necessary extension for future work.

In this chapter, the use of a numerical method for solving these equations for a

general case using the DSPM is described, so the rejection of various components

in a mixed solution containing multiple charged or uncharged compounds can be

predicted.

5.3 Model theory and solution details

5.3.1 The extended Nernst-Planck equation

The extended Nernst-Planck equation can be used to describe the transport of

solutes through a membrane [17, 148], and hence predict the rejection of a solute

by a membrane based on several physically meaningful parameters. It can be

written as shown in Eq. (5.1), where the terms on the right hand side represent

diffusion, convection, and in the case of a charged solute, electric migration,

respectively, for some solute i:

140
dc i z i c i K i , d D i d\
ji  K i , d Di  K i ,c c i J V  F (5.1)
dy RT dy

where ji is the solute flux, Di is the bulk diffusivity, zi is the electrical valence,

and ci is the concentration inside the membrane. The terms Ki,c and Ki,d represent

the hindered convection and diffusion of the solute inside the membrane,

respectively: the determination of these terms will be considered later. The overall

volumetric flux through the membrane is given by JV, and y is the axial coordinate

along the direction of the solution flux. The final term on the right hand side of

Eq. (5.1) representing transport due to electric migration depends on the Faraday

constant F, the ideal gas constant R, the absolute temperature T, and the electric

potential ȥ. The extended Nernst-Planck equation can thus predict the rejection

for uncharged solutes, charged solutes, or a mixture of the two.

5.3.2 Rejection of uncharged solutes

In the case of uncharged solutes, the electric migration term in Eq. (5.1)

disappears to yield the following:

dci
ji  K i ,d Di  K i ,c c i J V (5.2)
dy

It is assumed that the rejection of the uncharged solute can be modelled by a steric

partitioning coefficient ĭi which describes the accessibility of the pore to the

solute. The determination of the steric partitioning coefficient will be considered

141
later. This applies at the interface between the membrane and the external solution

both at the pore inlet (y = 0) and pore outlet (y = ǻy), such that:

ci , y 0 ĭi Ci ,W (5.3)

ci , y 'y ĭi Ci , p (5.4)

where ci is the concentration inside the membrane, Ci is the concentration of the

external solution, and the subscripts W and p represent the feed solution adjacent

to the membrane wall, and the permeate solution, respectively. The rejection of

solute i is defined as:

Ci , p
Ri 1 (5.5)
Ci ,W

By writing the flux of solute i as ji J V C i , p and integrating Eq. (5.2) along the

membrane pore between y = 0 and y = ǻy, the following solution can be easily

obtained:

ĭi K i ,c
Ri 1
1  1  ĭi K i ,c exp  Pei
(5.6)

Pei K i ,c J V ǻ\ K i ,d Di (5.7)

where Pei is the Peclet number.

142
The next step is the determination of the steric partitioning coefficient ĭ and

hindrance coefficients Ki,c and Ki,d. These differ depending on the assumed shape

of the solute, i.e. whether the solute is assumed to be spherical or non-spherical.

The simplest case is that of a spherical solute as follows.

5.3.2.1 Spherical solutes

The steric partitioning coefficient ĭi describes the rejection of the solute due to

steric (molecular sieving) effects. The simplest method for describing steric

effects is to assume a spherical solute, which is sound for small solutes. The

partitioning coefficient can then be written as a function of the membrane pore

radius rp and the radius of solute i, typically approximated as the Stokes radius,

ri,s:

ĭi 1  Oi 2 , Oi ri , s rp (5.8)

It should be noted that the use of the membrane pore radius need not necessarily

mean that well-defined pores physically exist within the membrane; instead, as

noted in [17], it may be regarded as a pore size which yields equivalent hindrance

to solutes as that of the polymeric structure of the membrane.

The hindrance factors for diffusion (Kd) and convection (Kc) describe how drag

forces inside the confined space of the pore hinder the diffusion and convection of

143
the solute [155], and are also functions of the ratio of the solute radius to pore

radius, Ȝ. Here, the expressions presented by Bowen et al. [148] have been used:

K i ,d 1.0  2.30Oi  1.154Oi  0.224Oi


2 3
(5.9)

K i ,c 2  ĭi 1.0  0.054Oi  0.988Oi 2  0.441Oi 3 (5.10)

5.3.2.2 Non-spherical solutes

For larger solutes, the shape of the solute plays an important role in the steric

interactions between the pore and solute. Several methods for determining the

steric partitioning coefficient for a non-spherical solute have been presented in the

literature based on representing the molecule in terms of a simplified geometric

shape. Example shapes include cylinders [156, 157], rectangular parallelipeds

[158-160], and prolate revolution ellipsoids [161]. The method described by Kiso

et al. [158-160] has been adopted in the present model, and is described briefly in

the following.

Approximating the solute as a rectangular parallelepiped, the solute length Li,S is

given by the distance between the two most distant points on the solute

(effectively the distance between the two most distant atoms in the molecule in

addition to their van der Waals radii). The line that joins these two most distant

points is defined as the L-axis. The solute width is determined by projecting the

solute particle onto the plane perpendicular to the L-axis, and constructing the

144
smallest possible rectangle on that plane which will fully enclose the projection of

the solute particle. Defining the area of this rectangle as Si, the solute width Wi,S

can then be calculated:

1
Wi , S Si (5.11)
2

As the solute is considered to be non-spherical, the solute orientation affects the

proportion of the pore area that is accessible to the solute. The length of the

projection of the solute against the pore surface is called the effective solute

length, and is given by:

Li , p Li , S cos D  2Wi , S sin D (5.12)

where the angle between the L-axis and the pore surface is Į. The steric partition

factor for some angle Į is thus given by:

2
§ ai ·
ĭi D ¨ ¸ (5.13)
¨r ¸
© p¹

2
§ Li , p ·
ai rp  ¨¨
2
¸¸  Wi , S (5.14)
© 2 ¹

A probabilistic approach is used to calculate an overall partition coefficient ĭi as

follows:

145
S /2
ĭi ³ ĭi D p D dD (5.15)
0

S /2
³ p D dD 1 (5.16)
0

where the probability function p(Į) represents the probability of an individual

soluWHEHLQJRULHQWHGDWDQDQJOHĮDQGPXVW satisfy Eq. (5.16). A common choice

of function is p(Į) = sin(Į).

A range of semi-empirical techniques are available to predict molecular geometry,

such as the MOPAC2009 software (Stewart Computational Chemistry, CO,

USA), which has been adopted for the compounds used in this study.

MOPAC2009 performs a series of optimisations to predict the most stable

conformation of a molecule, and returns a series of geometric parameters of the

optimised molecule, as well as Cartesian coordinates of each atom within the

molecule. Investigation of the MOPAC output indicated that the geometric

parameters output were sometimes unreliable, and hence direct processing of the

Cartesian output was required. For this purpose, a numerical routine was

developed to determine the solute length and width based on the Cartesian

coordinates of each atom and their associated van der Waals radii. The routine

employs a three-dimensional bounding box method to determine these geometric

parameters. The code for this routine and some sample output for several test

molecules (glucose, sucrose, cyanazine, glycerine, and Vitamin B12) is shown in

Appendix A.1.

146
Little information is available on determining hindrance factors for non-spherical

solutes. The approach adopted by Kiso et al. [159] uses the same expressions as

Eqs. (5.9) and (5.10); however, for the non-spherical solute the Stokes radius ri,s is

unknown. A correlation is thus used to approximate the Stokes radius (for use in

Eqs. (5.9) and (5.10) only) based on the solute width WS as follows:


ri , s u109 1.42 Wi , S u109  0.142 u10 9 (5.17)

5.3.2.3 Determination of membrane parameters for uncharged solutes

The only parameters thus needed to predict the rejection for a solution containing

uncharged solutes and an arbitrary membrane are the membrane pore radius rP

DQGSRUHOHQJWKǻy (sometimes given as ǻy /Ak, where Ak is the effective porosity

of the membrane). In principle the pore radius can be determined by inspection of

the membrane via methods such as atomic force microscopy [148]. However, in

practice, typically experimental measurement of the rejection of a single

uncharged solute is made, and rP and ǻy /Ak are then adjusted to achieve good

agreement with the experimentally measured rejection.

5.3.3 Rejection of charged solutes

For the case of charged solutes, the electric migration term is retained in Eq. (5.1),

which is no longer easily soluble by analytical means and must be solved

147
numerically. To do so, the electrical gradient through the pore, must be

determined, which for a system of n charged solutes is given by:

K i,c ci  Ci, p
n
zi J V
d\
¦Ki 1 i , d Di (5.18)
dy F n 2
RT i 1
¦ zi ci

This is subject to the condition of electroneutrality in the bulk solution:

¦zC
i 1
i i 0 (5.19)

and that electroneutrality is also upheld within the membrane pore:

¦z c
i 1
i i zX CX (5.20)

where zX and CX are the valence and density of the fixed membrane charge. It is

assumed this charge is constant throughout the membrane pore.

Eq. (5.1) is then solved numerically as a boundary value problem using the

following boundary conditions based on the assumption of Donnan equilibrium

with the solution considered sufficiently dilute that the activity coefficients are

unity:

§ zF ·
ci , y 0 ĭi Ci ,W exp¨  i '\ D , y 0 ¸ (5.21)
© RT ¹

148
§ zF ·
ci , y 'y ĭi C i , p exp¨  i '\ D , y 'y ¸ (5.22)
© RT ¹

where '\ D , y 0 and '\ D , y 'y are the Donnan potential at the pore inlet and outlet,

respectively. The steric partition coefficients ĭi are identical in meaning to the

uncharged case, and may be determined in the same manner, either using a

spherical model (in the case of a monatomic ion) or non-spherical model (in the

case of a polyatomic ion) as appropriate.

For all charged solutes at the pore inlet, the Donnan potential is identical, and the

same condition applies for the pore outlet. The concentrations just inside the pore

inlet ci,y=0 are determined according to this equality condition, and according to

the conditions specified in Eqs. (5.19) and (5.20), so that a set of n non-linear

equations is obtained as follows:

n n

¦ zi ci, y 0  z X C X  ¦ ziCi,W
i 1 i 1
0 (5.23)

1 §¨ C1,W ·¸ 1 §¨ Ci ,W ·¸
ln  ln 0, i 2,..., n (5.24)
zi ¨© c1, y 0 ¸¹ zi ¨© ci , y 0 ¸¹

For the simple case of n = 2 with univalent solutes, analytical expressions can be

derived relatively easily to solve for the concentrations ci,y=0. For n > 2 and / or

multivalent solutes, this becomes more difficult, and the equations tend to be

highly non-linear. A numerical method using a variant of Powell’s dogleg method

[162] is thus used to obtain the concentrations for the general case.

149
Determining the outlet boundary condition presents an additional difficulty as the

permeate concentration is not yet known. This is overcome by writing the

permeate concentration for each solute in terms of the feed concentration and the

rejection, viz.

Ci , p 1 Ri Ci ,W (5.25)

so using Eq. (5.22) one can write:

§ zF ·
ci , y 'y ĭi 1  Ri Ci ,W exp¨  i '\ D , y 'y ¸ (5.26)
© RT ¹

It is now possible to solve the boundary value problem consisting of the n

ordinary differential equations defined by Eq. (5.1), and the 2n boundary

conditions defined by Eqs. (5.21) and (5.26). This is done numerically with the

rejections Ri as a set of additional unknown parameters that also must be solved

for in order to establish the Donnan potential at the pore outlet. Implementations

of this numerical solution have been written in both MATLAB and Fortran 90,

using the boundary value problem solvers developed for these languages by

Shampine et al. [163, 164]. The Donnan potential at the pore outlet is determined

within the boundary value problem solver by iteratively solving the system of

non-linear equations described by Eqs. (5.23) and (5.24).

150
5.3.3.1 Determination of membrane parameters for charged solutes

In addition to the parameters rP and ǻy /Ak used for uncharged solutes, the

additional parameters describing the charge of the membrane, zX and CX must be

determined for charged solutes. For virtually all practical membranes, the inherent

charge of the membrane is negative in sign, so zX = -1, though the numerical

implementation can handle either case. To determine CX, it should be possible to

perform first perform rejection experiments for uncharged solutes to obtain rP and

ǻy /Ak, and then perform additional rejection experiments for a single salt solution

and adjust CX to obtain a good fit. In the literature however, there are often

substantial differences between ǻy /Ak values between the uncharged and charged

cases [141, 148]. The reasons for this discrepancy are not fully clear, as discussed

later.

In addition, it has also been observed [17, 141, 147-149] that the charge density

CX also varies with the feed concentration of the salt solution. This can be

expressed in terms of an isotherm [17], which can be determined by rejection

experiments for a single salt solution at varying feed concentrations.

5.3.4 Rejection of mixtures of charged and uncharged solutes

Combining the approaches described previously to determine the rejections for a

solution containing charged and uncharged solutes is straightforward.

Calculations for any uncharged solutes which depend only on steric effects can be

151
done independently of the calculations for charged solutes. It is probable however

that many solutes may exhibit polar effects which may further complicate the

solute-solute and solute-membrane interactions. These interactions are not

considered here. A flowchart illustrating the combined procedure is shown in

Figure 5.1.

152
Figure 5.1 – Flowchart of calculation procedure for combination of charged and uncharged solutes

153
5.3.5 A note on concentration polarisation and volumetric flux

It is important to note that the model described here does not include the effect of

the ubiquitous phenomenon of concentration polarisation. This omission is

deliberate, as this numerical routine has been developed for the purpose of

inclusion in a coupled hydrodynamic and pore model using ANSYS CFX, as will

be discussed in Chapter 6. At this stage it is sufficient to note that the

concentration polarisation effects may be satisfactorily modelled outside of the

pore model (i.e. within a hydrodynamic model of the bulk flow in the membrane

channel) provided the hydrodynamic model maintains the electroneutrality

condition within the concentration polarisation boundary layer. In this way, the

conditions at the pore inlet can be obtained from the hydrodynamic solution

including the effects of concentration polarisation. The model presented here thus

predicts the real rejection of the membrane. It is emphasised all comparisons to

experimental data in this chapter are presented in terms of real rejection, which

have been obtained by other authors from observed rejections using correlations

for mass transfer coefficients [e.g. 17, 141, 148].

It is also assumed that the permeate volumetric flux JV is known, which can be

easily measured experimentally. For use in a predictive model of the membrane

filtration process using the coupled hydrodynamic / pore model approach, the flux

JV can be determined using a Darcy’s law formulation if the hydrodynamic

permeability of the membrane is known, or equivalently, a Hagen-Poiseuille

formulation.

154
5.4 Computational and numerical considerations

5.4.1 Use of existing numerical routines and libraries

As discussed in Section 5.3.3, numerical codes to allow these rejection

calculations have been developed both in MATLAB and Fortran 90, using already

available numerical routines where available. The use of robust public domain

numerical routines for general purpose calculations saves considerable time in

developing these codes, and gives confidence in the numerical predictions due to

the considerable testing and verification of such codes. A summary of the routines

used in the numerical codes is shown in Table 5.1.

Table 5.1 – Summary of existing numerical libraries and routines used in development of predictive rejection

model code

Routine / library name and language Details

fsolve (MATLAB) [165] Solvers for systems of non-linear equations. Used to solve
hybrd1 (Fortran 90, part of minpack Donnan partitioning problem for charged solutes at pore inlet and
library) [166] outlet (see Eqns. (5.18) through (5.26)).

quad (MATLAB) [165]


Numerical quadrature (integration) codes. Used to perform
qag (Fortran 90, part of quadpack
numerical integration to determine overall steric partitioning
library) [166]
coefficient for non-spherical solutes (see Eqn. (5.15)).

Boundary value problem solvers. Used to solve extended Nernst-


bvp4c (MATLAB) [164] Planck equations (see Eqns. (5.1), (5.21), and (5.26)) to determine
BVP_SOLVER(Fortran 90) [163] concentration profile throughout membrane pores and hence
determine solute rejections as unknown parameters.

155
5.4.2 Program structure

Some illustrative sample code for the MATLAB version of the numerical method

is shown in Appendix A. Briefly, there are five main routines:

x molprjn2 – this routine determines the molecular length and width for a

non-spherical solute based on the Cartesian coordinate output from the

MOPAC2009 software;

x kiso – this routine determines steric partitioning coefficients for non-

spherical solutes;

x doneqn3 – this routine calculates the Donnan partitioning equilibrium

constants for an arbitrary number of charged solutes so that the Donnan

potential for each solute can be calculated;

x np3n – this routine solves the extended Nernst-Planck equations for an

arbitrary number of charged solutes; and

x unchargedJv – this routine calculates the rejection for an arbitrary number

of uncharged solutes by solving a simplified version of the extended

Nernst-Planck equations.

There are several techniques of interest contained within these routines which will

be discussed briefly in the following.

156
5.4.3 Determination of Donnan partitioning equilibrium constants

The determination of the Donnan equilibrium constants for more than three

charged solutes must be done numerically, using existing routines to solve the

system of non-linear equations described by Eqns. (5.23) and (5.24). The most

important step to ensure these solvers work efficiently and produce the correct

solution is to provide an accurate initial estimation of the solution, which in this

case is the concentration of each solute inside the membrane pore. A reasonably

accurate estimation is made with the knowledge that the pore concentration of

counter-ions (solutes with the opposite charge to the membrane) is typically much

higher than the concentration in the bulk solution. Similarly, the concentration of

co-ions inside the pore is typically lower than that in the bulk solution. The initial

estimate of the concentration is thus given by:

zi zi
ci , y Ci , y CX , 0 (counter - ion) (5.27)
zX zX

zi 1 zi
ci , y Ci , y , !0 (co - ion) (5.28)
zX CX zX

Using this estimate was found to result in rapid convergence to a solution.

Confirmation that the true solution has been found can be obtained by using the

equilibrium constants to calculate the vector of Donnan potentials for the system,

which should be identical for every ion. The numerical routine uses this as a

check to ensure the correct solution has been found.

157
In a small proportion of cases, particularly where the membrane charge density CX

is large, the system of equations becomes close to singular, and the solver may

return a solution which yields a Donnan potential which is not equal for all ions.

In these cases, it is very difficult to provide a better estimate at the solution than

given by Eqns. (5.27) and (5.28). In this case, a crude but effective method is used

whereby the initial estimate at the solution is multiplied by a random number, and

then used as an updated initial estimate. The system of equations is again solved,

and the equality of the Donnan potential is checked. This process is repeated until

the Donnan potential equality condition is satisfied. In practice, the true solution

tends to be found within several iterations of this random multiplication method,

which is relatively small in terms of computational expense.

5.5 Results and discussion

A set of test cases for the model has been derived from experimental results

reported in the literature as follows: for rejection of uncharged compounds and

single salt solutions, the results of Bowen and Mohammad [141] have been used;

for rejection of single and multiple salt solutions, the results of Bowen and

Mukhtar [17] have been used. Finally, to demonstrate the general applicability of

the method, a theoretical investigation of the rejection of complex mixtures

containing four or more charged solutes has been undertaken.

158
5.5.1 Data from Bowen and Mohammad

5.5.1.1 Rejection of uncharged solutes

This study [141] used a CA30 cellulose acetate membrane (Hoescht Separation

Products, Germany) to investigate the rejection of a dye / salt solution. To

characterise the membrane pore radius using a spherical pore model, a series of

experiments was first performed with several uncharged compounds (glycerin,

glucose, sucrose, and Vitamin B12). These experimental results are used here to

validate the present model and present an improved characterisation of the CA30

membrane using the non-spherical pore model described previously.

The geometry of the uncharged compounds was determined using the

MOPAC2009 software (Stewart Computational Chemistry, CO, USA) using the

PM6 parameter optimisation [167]. The molecular length and width were then

determined by post-processing of the Cartesian coordinates generated by

MOPAC2009 and the associated van der Waal’s radius for each atom to obtain

the necessary molecular projections. The determined geometric parameters, as

well as the properties previously determined in [141] are shown in Table 5.2.

159
Table 5.2 - Physical properties for uncharged solutes used in [141]

Solute ri,S (nm) * Di × 109 (m2 s-1) * Li,S (nm) Wi,S (nm)

Glycerin 0.260 0.95 0.827 0.350

Glucose 0.365 0.69 1.067 0.441

Sucrose 0.471 0.52 1.138 0.515

Vitamin B12 0.740 0.33 2.244 1.109

* From [141].

In their characterisation of the membrane, Bowen and Mohammad [141] used

both the pore radius rp and the thickness-SRURVLW\ SDUDPHWHU ǻy/Ak as fitting

parameters, i.e. separate rp DQGǻy/Ak were fitted for each compound. An averaged

value of the rp was then adopted as the effective pore radius for subsequent

analysis with salt solutions. Clearly this is not ideal, as one might expect rp and

ǻy/Ak to be consistent regardless of the compound, as they are properties of the

membrane. Using a numerical fitting routine with a non-linear least-squares

algorithm, the analysis was undertaken here using a non-spherical pore model.

The results of the fitting are shown in Table 5.3, along with the previous fitting

results, and the membrane rejections as a function of flux are shown in Figure 5.2.

160
Table 5.3 – Fitting results for uncharged solutes used in [141]

Solute Previous fit * Current – non-spherical

rp (nm) ǻy/Ak (μm) rp (nm) ǻy/Ak (μm)

Glycerin 0.62 2.65 1.11 4.11

Glucose 0.61 1.92 1.11 4.11

Sucrose 0.66 0.74 1.11 4.11

Vitamin B12 0.87 0.39 1.48† 4.11†

* From [141] using spherical pore model.



Vitamin B12 pore radius fitted separately from other solutes.

100

90

80

70

60
Ri (%)

50
Bowen and Mohammad (1998)
40 Non-spherical pore model
Gly cerin
30 Glucose
Sucrose
Vitamin B1 2
20

10

0
0 5 10 15
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 5.2 - Comparison of experimental rejection data [141] and numerically computed rejections as a

function of flux

161
A reasonably good fit is obtained for glycerine, glucose and sucrose using fixed rp

DQG ǻy/Ak. This gives confidence that the use of a non-spherical pore model is

able to accurately characterise the membrane in terms of actual physical

parameters. It was found that it was difficult to fit the data for Vitamin B12 well

due to its large molecular size; this was dealt with in [141] by fitting a much

larger pore radius of 0.87 nm for Vitamin B12. Here, Vitamin B12 has been again

been treated separately for the sake of completeness with the same fixed ǻy/Ak as

the other solutes to obtain a larger pore radius of 1.48 nm. This should not be

interpreted as meaning that the pore radius would vary by solute. Instead, as noted

in [141], transmission of Vitamin B12 molecules probably occurs through a small

proportion of larger pores which are neglected in the current analysis by assuming

a constant pore size.

It can also be noted that the pore radius determined using the non-spherical pore

model (1.11 nm) is significantly larger than the averaged representative pore

radius of 0.63 nm previously determined [141] using a spherical pore model. This

is consistent with previous comparisons using both spherical and non-spherical

pore models [160], and the ratio of the spherical radius to the non-spherical radius

is also consistent with these results. As well, the ratio obtained for Vitamin B12

(1.48 nm : 0.87 nm) is similar.

162
5.5.1.2 Rejection of single salt solutions

In [141], solutions of sodium chloride, lithium chloride, and potassium chloride

were filtered using the same CA30 membrane to characterise the membrane in

WHUPVRIǻy/Ak and the membrane charge CX, using the representative rp = 0.63

QPREWDLQHGIURPDVSKHULFDOSRUHPRGHO,QWKLVFKDUDFWHULVDWLRQǻy/Ak differed

significantly from that obtained with uncharged solutes. As this should in theory

be a fixed physical property of the membrane, this is again somewhat less than

satisfactory.

Here, this data is revisited to characterise the membrane charge CX, using a

consistent rp DQGǻy/Ak from the characterisation using uncharged solutes with a

non-spherical pore model described in Section 5.5.1.1. For this particular case, all

the ionic solutes are monatomic, so the characterisation of the solute shapes as

spherical and hence the determination of the steric partitioning coefficient as per

Eq. (5.8) is justified. It is stressed however that the pore radius obtained using a

non-spherical pore model (1.11 nm) is used throughout so as to present a

consistent method for mixtures of monatomic (spherical) and polyatomic (non-

spherical) solutes. The physical properties of the ionic solutes are shown in Table

5.4.

163
Table 5.4 - Physical properties for ionic solutes used in [141]

Ion ri,S (nm) Di × 109 (m2 s-1)

Na+ 0.184 1.33

Cl- 0.121 2.03

K+ 0.125 1.95

Li+ 0.238 1.03

The model was first fitted in terms of a constant CX to rejection measurements for

the three salts at a fixed bulk concentration of 10 mol m-3, as shown in Figure 5.3.

A reasonably good fit was obtained with CX = 98.4 mol m-3 using the same

parameters as for the uncharged solute fitting (rp =  QP ǻy/Ak = 4.11 μm).

The fitted membrane charge is significantly larger than that in [141], which is to

be expected as the pore radius is also significantly larger.

The model was then fitted to rejection data for sodium chloride at varying bulk

concentrations. In this case, the membrane charge can reasonably be expected to

vary with the ionic strength of the solution. Using the same procedure and rp and

ǻy/Ak values, the results obtained from the model fitting are shown in Table 5.5

and Figure 5.4.A reasonably good fit is again obtained using a consistent set of

parameters between the uncharged and charged solute cases. As expected, the

fitted CX are larger than the previous fit due to the larger pore radius. The

variation in CX with the bulk concentration is shown in Figure 5.5, and was found

164
to fit a linear isotherm well, as in [141]. The isotherm for the current fitted

parameters is given by:

CX 8.23CB  24.06 (5.29)

Table 5.5 – Fitting results for single salt solutions used in [141]

Previous fit Current fit


CB
Salts (mol m- rp ǻy/Ak CX rp ǻy/Ak CX
3
)* (nm) (μm) (mol m- (nm) (μm) (mol m-
3 3
) )

NaCl, KCl, 10 0.63 6.2 39.2 1.11 4.11 98.4


LiCl

NaCl 0.5 0.63 6.2 10.7 1.11 4.11 27.6

1.0 0.63 6.2 11.4 1.11 4.11† 29.3

5.0 0.63 6.2 23.6 1.11 4.11† 59.4

50.0 0.63 6.2 184.1 1.11 4.11† 468

100.0 0.63 6.2 333.6 1.11 4.11† 832

* Bulk concentration - equivalent to the wall concentration Ci,W assuming no concentration polarisation.

165
20

18

16

14

12
Ri (%)

10

6 KCl
LiCl
NaCl
4 Fitted model

2 4 6 8 10 12
Flux Jv × 106 (m3 m-2 s -1)

Figure 5.3 - Comparison of experimental rejection data [141] and numerically computed rejections as a
function of flux for single salt solutions at fixed concentration

0.5 mol m -3
1.0 mol m -3
-3
60 5.0 mol m
10 mol m -3
50 mol m -3
-3
100 mol m
50 Fitted model

40
Ri (%)

30

20

10

2 4 6 8 10 12
Flux Jv × 106 (m3 m-2 s -1)

Figure 5.4 - Comparison of experimental rejection data [141] and numerically computed rejections as a
function of flux for sodium chloride at varying concentrations

166
1000

900

800

700

600
CX (mol m-3)

500

400

300 Fitted data points


Fitted isotherm

200

100

0 20 40 60 80 100 120
-3
Bulk concentration CB (mol m )

Figure 5.5 – Variation of charge density as a function of bulk concentration using experimental data from

[141]

5.5.2 Data from Bowen and Mukhtar

This study [17] differs from the previously mentioned study [141] as different salt

solutions were used to characterise a NF membrane in terms of rpǻy/Ak, and CX,

without initial use of uncharged solutes to determine rp. The same approach is

applied here to demonstrate the consistency of the computational method.

167
5.5.2.1 Rejection of single salt solutions

In this study [17], six different NF membranes were used in a stirred cell

arrangement and rejections were experimentally measured for solutions of sodium

chloride and sodium sulphate. These observed rejections were converted to real

rejections (as used throughout here) by use of a mass transfer coefficient to

describe the effects of concentration polarisation. The membranes were initially

characterised by assuming a uniform pore radius of 1 nm and adjusting CX and

ǻy/Ak to obtain a good fit between the experimental results and their solutions to

the extended Nernst-Planck equations shown in Eq. (5.1). It should be noted that

in all simulations it was assumed by Bowen and Mukhtar [17] that steric effects

were not significant (i.e. the steric partitioning coefficients are unity), and that all

membranes have a negative effective charge (zX = -1). A slightly different

expression was also used to calculate the hindrance coefficients Ki,c and Ki,d;

however, the expressions given by Eqs. (5.9) and (5.10) have been retained here

as they are applicable for a wider range of Ȝ. Relevant physical parameters for the

sodium, chloride and sulphate ions are shown in Table 5.6, including molecular

length and width parameters derived for the polyatomic sulphate ion.

168
Table 5.6 - Physical properties for solutes from [17]

Ion ri,S (nm) * Di × 109 (m2 s-1) * Li,S (nm) Wi,S (nm)

Na+ 0.1840 1.333 - -

Cl- 0.1207 2.031 - -

SO42- 0.2309 1.062 0.566 0.271

* From [17].

Using the same approach, the present solution method was applied to model the

experimental results presented. A comparison of the experimental results of

Bowen and Mukhtar for a single membrane (the PES5 membrane) with a 10-3 M

solution of sodium chloride [17] with the rejection predicted by the present

method, both with and without steric effects, is shown in Figure 5.6, and the

resulting fitted parameters are shown in Table 5.7. Again, the steric partitioning

coefficients were determined for a spherical solute as the ions are monatomic. The

parameters were adjusted by way of a numerical fitting routine using a non-linear

least-squares algorithm. A slightly improved fit with the experimental data is

observed compared to the parameters obtained previously in [17], due to the

different expressions for the hindrance coefficients.

169
30

25

20
Ri (%)

15

10
Numerical - no steric effects
Numerical - with steric effects
5 Numerical - parameters from Bowen and Mukhtar (1996)
Experimental

5 10 15 20 25 30 35 40
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 5.6 - Comparison of experimental rejection data from Bowen and Mukhtar [17] and numerically

computed rejections for PES5 membrane for 10-3 M sodium chloride solution using ǻy/Ak and CX as fitting

parameters

Table 5.7 - Comparison of fitting for ǻy/Ak and CX for PES5 membrane used by Bowen and Mukhtar [17] for

10-3 M sodium chloride solution

Configuration ǻy/Ak (×10-5 m) CX (eq m-3) rp (nm) *

No steric effects 1.92 4.35 1

Steric effects – spherical pore model 1.47 2.85 1

Parameters from [4] – no steric effects 2.00 4.39 1

* Assumed value – not fitted.

170
The results in Table 5.7 show significant differences in the fitted parameters

depending on whether steric effects are included, indicating that steric interactions

may be important in this system. It is also possible to further characterise the

membrane in terms of the pore radius, which was previously assumed to be 1 nm.

Allowing the pore radius to be fitted as an additional parameter using the least-

squares algorithm yields the fitting parameters shown in Table 5.8. A graph of the

fit is shown in Figure 5.7.

30

25

20
Ri (%)

15

Numerical - no steric effects


10 Numerical - with steric effects
Experimental

5 10 15 20 25 30 35 40
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 5.7 - Comparison of experimental rejection data from Bowen and Mukhtar [17] and numerically

computed rejections for PES5 membrane for 10-3 M sodium chloride solution using ǻy/Ak, CX, and rp as

fitting parameters

171
Table 5.8 - Comparison of fitting IRUǻy/Ak, CX, and rp for PES5 membrane used by Bowen and Mukhtar [17]

for 10-3 M sodium chloride solution

Configuration ǻy/Ak (×10-5 m) CX (eq m-3) rp (nm)

No steric effects 1.96 4.31 1.01

Steric effects – spherical pore model 1.60 3.00 1.14

Bowen and Mukhtar [17] then fitted experimental data for sodium sulphate at a

feed concentration of 3×10-3 M, and observed a larger dependence of the rejection

behaviour with the pore radius due to the larger size of the sodium sulphate

molecule. They obtained a pore radius of 0.72 nm, which was then used to refit

the data for sodium chloride at a concentration of 10-3 M. Here, the data for

sodium sulphate (3×10-3 M) is fitted with the same rp DQG ǻy/Ak as for sodium

chloride (10-3 M), which should be consistent for both solutions. However, the

charge density CX is allowed to vary, as it is expected to be a function of ionic

strength. Additionally, the non-spherical pore model can also be used for the

sulphate ion with the characteristic length and width given in Table 5.6. From this

point onwards the combination of the spherical and non-spherical pore models is

referred to simply as the non-spherical pore model. The variation in rejection is

shown in Figure 5.8, with the parameters obtained shown in Table 5.9.

172
100

90

80

70

60
Ri (%)

No steric effects
50 Steric effects - spherical
Steric effects - non-spherical
Experimental
40

30

20

10

0
0 5 10 15 20 25
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 5.8 - Comparison of experimental rejection data from Bowen and Mukhtar [17] and numerically

computed rejections for PES5 membrane for 3×10-3 M sodium sulphate solution using CX as fitting parameter

Table 5.9 - Comparison of fitting for CX for PES5 membrane used by Bowen and Mukhtar [17] for 3×10-3 M

sodium sulphate solution

Configuration ǻy/Ak (×10-5 m) rp (nm) CX (eq m-3)

No steric effects 1.96 1.01 22.0

Steric effects – spherical pore model 1.60 1.14 13.8

Steric effects – non-spherical pore model 1.60 1.14 16.5

173
Several comments can be made regarding the fitting of the sodium sulphate

solution: first, a good fit is possible using the same pore radius as obtained in

Figure 5.7 for sodium chloride both with and without the inclusion of steric

effects (1.01 and 1.14 nm, respectively). This is in contrast with the results in

[17], where a significantly smaller pore radius (0.72 nm) was required to fit the

sodium sulphate data. Secondly, using a non-spherical pore model for the sulphate

ions gives very similar predicted rejections as for a spherical pore model with

consistent pore radius and membrane thickness parameters. This indicates that use

of a combination of spherical and non-spherical pore models is appropriate when

such mixtures occur, and shows the internal consistency of the approach treating

the pore radius and membrane thickness as fixed physical parameters.

Bowen and Mukhtar [17] also measured the rejections of both salts at various

concentrations for a fixed flux to determine how CX varied with concentration.

Using the same data, this procedure was repeated to obtain two isotherms

describing the variance in CX with the concentration when expressed in units of

equivalents per cubic metre: one isotherm using the spherical pore model, and the

other using the combined spherical and non-spherical pore models as described

above. The resulting isotherms are shown in Figure 5.9, and as per [17] are of the

type:

log10 C X s log10 C B  q (5.30)

174
1.8

1.6 Spherical model


Non-spherical model
Spherical model isotherm

1.4 Non-spherical model isotherm

1.2
Log10 of CX (mol m-3)

0.8

0.6

0.4

0.2

0
-1 0 1 2
Log10 of bulk concentration CB (mol m-3)

Figure 5.9 - Variation of charge density as a function of bulk concentration using experimental data from [17]

The fitted parameters s and q are shown in Table 5.10. This allows for the

prediction of the rejection of mixed salt solutions of arbitrary ionic strength,

which is addressed in the following.

Table 5.10 – Fitted parameters for charge density isotherms using experimental data from [17]

Configuration s q

Steric effects – spherical pore model 0.509 0.676

Steric effects – non-spherical pore model 0.523 0.699

175
5.5.2.2 Rejection of multiple salt solutions

Bowen and Mukhtar [17] measured the rejection of mixtures of sodium chloride

and sodium sulphate at various molar ratios, and then predicted the rejection for

each ion using an isotherm to describe the charge density CX for each molar ratio.

This is an interesting test for the model as the experimental data display widely

variable rejection behaviour, including the phenomenon of negative rejection.

The current method has been applied to reproduce these results. For brevity, the

results presented here are only for the case where the combined spherical and non-

spherical model is used (depending on whether the ions are monatomic or

polyatomic). Using the membrane parameters described in Table 5.9 and Table

5.10 for the non-spherical pore model (rp = 1.14 nm, ǻy/Ak = 1.60 × 10-5 m, s =

0.523, q = 0.699), the resulting model predictions are shown in Figure 5.10 for

molar ratios of 0.2, 0.4, 0.6 and 0.8 (NaCl:Na2SO4). Very good agreement is

observed between the numerical model and the experimental data, including the

negative rejection of the Cl- co-ion.

176
100 100
(a) (b)
80 80

60 60

40 40

20 20
Ri (%)

Ri (%)
0 0

-20 -20

-40 -40

-60 -60

0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

100 100
(c) (d)
80 80

60 60

40 40

20 20
Ri (%)

Ri (%)

0 0 +
Na model
-
Cl model
-20 -20 SO2- model
4
+
Na experimental
-40 -40 -
Cl experimental
2-
SO4 experimental
-60 -60

0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

Figure 5.10 - Comparison of experimental rejection data from Bowen and Mukhtar [17] and numerically

computed rejections for PES5 membrane for 1×10-3 M salt solution using membrane charge density isotherm

for non-spherical pore model with molar ratios of: (a) 0.2, (b) 0.4, (c) 0.6, (d) 0.8 (NaCl:Na2SO4)

177
5.5.3 Predictions for solutions with four or more charged solutes

The model presented above is in theory capable of predicting the rejection for a

mixture containing arbitrary numbers of charged and uncharged solutes. Presented

in the following are some predictions for the rejection behaviour for solutions

containing four or more charged solutes using the combined spherical and non-

spherical pore models. It should be noted that these predictions are for theoretical

interest only as little detailed experimental data is available for such complex

mixtures.

The first solution to be modelled is a 1×10-3 M salt solution of sodium chloride,

sodium sulphate, and calcium chloride containing four different charged solutes

(Na+, Cl-, SO42-, and Ca2+). The membrane parameters used, including the charge

density isotherm, are identical to that in Figure 5.10, with four different molar

ratios (NaCl:Na2SO4:CaCl2) of 0.333:0.333:0.333, 0.25:0.25:0.5, 0.25:0.5:0.25,

and 0.5:0.25:0.25. The predicted rejections are shown in Figure 5.11. The

rejection curves are broadly similar to that observed for the three-ion case (Figure

5.10), with the addition of the Ca2+ counter-ion (rp = 0.236 nm, Di = 0.236 × 109

m2 s-1), which experiences greater rejection than the Na+ counter-ion due to its

greater valence and larger ionic radius. Negative rejection is still predicted for the

Cl- co-ion at low fluxes in all cases.

178
100 100
(a) (b)
80 80

60 60

40 40

20 20
Ri (%)

Ri (%)
0 0

-20 -20

-40 -40

-60 -60

0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

100 100
(c) (d)
80 80

60 60

40 40

20 20
Ri (%)

Ri (%)

0 0

-20 -20 Na model


+

-
Cl model
-40 -40 SO model
2-
4
2+
Ca model
-60 -60

0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

Figure 5.11 - Numerically computed rejections for PES5 membrane for 1×10-3 M salt solution using

membrane charge density isotherm for non-spherical pore model with molar ratios of: (a) 0.333:0.333:0.333,

(b) 0.25:0.25:0.5, (c) 0.25:0.5:0.25, (d) 0.5:0.25:0.25 (NaCl:Na2SO4:CaCl2)

179
A final demonstration of the model has been made for a solution comprising

sodium chloride, sodium sulphate, calcium chloride, potassium chloride and

lithium chloride. This solution therefore contains six separate ions (Na+, Cl-, SO42-

, Ca2+, K+, and Li+). Simulations were performed for a 1×10-3 M salt solution

using identical membrane parameters including the charge density isotherm as for

Figure 5.10 and Figure 5.11, using the molar ratios listed in Table 5.11. The

predicted rejections are shown in Figure 5.12. Again, some negative rejection is

predicted at low flux for the Cl- co-ion, with positive rejections predicted for the

SO42- co-ion and all counter-ions.

Table 5.11 – Molar ratios used for simulation of six ion 1×10-3 M salt solution

Molar fraction
Configuration
NaCl Na2SO4 CaCl2 KCl LiCl

(a) 0.2 0.2 0.2 0.2 0.2

(b) 0.4 0.15 0.15 0.15 0.15

(c) 0.15 0.4 0.15 0.15 0.15

(d) 0.15 0.15 0.4 0.15 0.15

(e) 0.15 0.15 0.15 0.4 0.15

(f) 0.15 0.15 0.15 0.15 0.4

180
100 100
(a) (b)
80 80

60 60
Ri (%)

Ri (%)
40 40

20 20

0 0

-20 -20
0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

100 100
(c) (d)
80 80

60 60
Ri (%)

Ri (%)

40 40

20 20

0 0

-20 -20
0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

100 100
(e) (f)
Na + model
80 80 Cl- model
SO24 - model
60 60 Ca
2+
model
K+ model
Ri (%)

Ri (%)

Li+ model
40 40

20 20

0 0

-20 -20
0 10 20 30 40 0 10 20 30 40
Flux Jv × 106 (m3 m-2 s -1) Flux Jv × 106 (m3 m-2 s -1)

Figure 5.12 - Numerically computed rejections for PES5 membrane for 1×10-3 M salt solution using

membrane charge density isotherm for non-spherical pore model with molar ratios as listed in Table 5.11

181
5.6 Conclusions

The use of the extended Nernst-Planck equation in conjunction with steric pore

models for both spherical and non-spherical solutes provides a general method for

predicting the rejection performance of nanofiltration (as well as reverse osmosis)

membranes for both charged and uncharged solutes. The analysis of several sets

of experimental data from the literature showed that combining spherical and non-

spherical pore models to predict steric hindrance allows for membranes to be

characterised and modelled using a consistent set of physical parameters (namely,

the membrane pore radius rp and the membrane thickness parameter ǻy/Ak). This

allows for accurate prediction of solutions containing arbitrary mixtures of

uncharged (non-spherical) solutes, and charged solutes, both monatomic

(spherical) and polyatomic (non-spherical). The fact that good agreement can be

achieved with experimental data using consistent membrane parameters for

different solutes gives good confidence that the chosen physical parameters are

indeed a good choice for membrane characterisation. The use of isotherms to

describe the variation in membrane charge density CX in terms of the bulk

concentration (expressed in units of equivalents per cubic metre) also has been

shown to provide a robust method of predicting rejection of multiple ion

solutions.

Additionally, the use of a numerical solution to the Donnan partitioning model for

charged solutes allows for prediction of the rejection for solutions containing

182
large numbers of charged solutes with the extended Nernst-Planck equations,

which to the authors knowledge has not been described in the literature for

mixtures of more than four unique solutes. This can easily be extended to very

complex mixtures of charged and uncharged solutes as the uncharged solute

rejections can be predicted independently of those of the charged solutes.

It is also acknowledged that the DSPM, of which the present model is a small

refinement, has some limitations which can be potentially addressed by the further

inclusion of dielectric exclusion effects (i.e. the DSPM-DE). However, the

numerical techniques used for solving the equations of the present model are

directly applicable to the DSPM-DE equations, and the future extension of the

model to include these effects is a feasible and worthwhile avenue of study.

Finally, this method is suitable to be incorporated into a coupled hydrodynamic

and pore model, so that the entire predictive model of a NF or RO membrane

system may be reasonably constructed. The development of this coupled model

approach is described in Chapter 6.

183
Chapter 6

A fully predictive membrane filtration model: coupled

hydrodynamic and predictive rejection modelling

6.1 Introduction

The models described in Chapter 4 have been successful in predicting the

hydrodynamic behaviour and growth of the concentration boundary layer within a

membrane channel. However, as true predictive models they are not entirely

useful, as the permeate flux through the membrane wall is a function of the

osmotic pressure, which depends on the rejection factor of a particular solute for a

particular membrane at a particular transmembrane pressure. In all the cases

modelled so far, this rejection factor has been provided based on experimental

observations of physical membrane systems.

It would seemingly be extremely useful to be able to predict the rejection of an

arbitrary solute or combination of solutes for an arbitrary membrane without

resorting to experimental studies, and hence provide a generalised computational

model for the membrane filtration process. The numerical method for

characterisation of membranes and prediction of membrane rejection presented in

Chapter 5 allows for this aim to be achieved when coupled with a rigorous

184
hydrodynamic model, at least for NF and RO systems (and potentially some UF

systems).

This chapter describes the development of a general model which couples the

hydrodynamic models described in Chapter 4, and the predictive rejection models

described in Chapter 5, to describe the membrane filtration process in

considerable detail, and for as general a case as feasible.

6.2 Coupled modelling of solutions containing uncharged solutes

For uncharged solutes, analytical solutions to the Nernst-Planck equations are

available (as described in Section 5.3.2), and hence predicting the rejection is

fairly straightforward in computational terms. Because of this, it is possible to

perform all of the calculations for a coupled hydrodynamic-predictive rejection

model for uncharged solutes entirely within the CFX environment, using the

native Command Expression Language (CEL) to perform all calculations. The

following section describes the development of a CEL algorithm for coupling the

uncharged solute rejection model with the hydrodynamic model.

6.2.1 Model details

As described in Section 5.3.2, the prediction of uncharged solute rejections is

fairly simple. The most involved numerical procedure in the method is the

numerical integration to find the overall steric partitioning coefficient in the case

185
of a non-spherical solute, as in Eq. (5.15). This can be achieved relatively simply

using native CEL expressions within CFX, using a Simpson’s rule quadrature

method. CEL expressions are then calculated for the rejection at each

computational point on the membrane surface, using local values obtained from

the hydrodynamic solution. A flowchart illustrating the computational procedure

for the coupled model with uncharged solutes is shown in Figure 6.1.

186
Figure 6.1 – Flowchart of computation procedure for coupled hydrodynamic-mass transfer model for

uncharged solutes only

187
6.2.2 Model verification

Using the coupled model, verification was undertaken using the experimental

conditions measured by Geraldes et al. [94] for two uncharged compounds

(sucrose and PEG1000). All simulation conditions were identical to that used in

Section 4.2.5, except that now with the coupled rejection model, the solute

rejection is calculated within the hydrodynamic model based on the local

hydrodynamic conditions, rather than being explicitly specified. In addition, this

approach allows for the solute rejection to vary spatially over the computational

domain (i.e. over the membrane surface) depending on the local hydrodynamic

conditions.

The MOPAC2009 software was used to predict the solute molecule geometry for

the two test compounds as shown in Table 6.1. For all simulations, no osmotic

pressure correction factor was used, in order to properly elucidate the effect of the

spatially variable rejection model on the predicted permeate fluxes, if any.

Table 6.1 - Molecular properties for uncharged test compounds derived using MOPAC2009 software

Molecular
Solute length LS Solute width WS
Compound Molecular formula weight
(nm) (nm)
(g mol-1)

Sucrose C12H22O11 342.30 0.973 0.480

PEG1000 H-(OCH2CH2)n-OH* ~ 1000 5.459 1.549

* Generic formula for polyethylene glycol – here n = 22 to correspond with listed average molecular weight of 1000 g mol-1

188
Detailed information on the geometric properties of the CDNF50l membrane was

unavailable; however de Pinho et al. [168] performed a series of experiments in

which the average pore radius for this membrane was characterised as 0.52 nm.

This value was obtained by comparing results from a similar pore model,

assuming a spherical molecule, and experimental studies. Using a similar

approach, a pore radius of 0.79 nm gave good agreement with the experimentally

measured rejection factors for the present non-spherical model. As the rejection in

the current formulation can vary axially with the local osmotic pressure and TMP,

which depends on the hydrodynamic solution, a representative rejection value

averaged over the membrane surface was compared against the measured

rejection in order to determine the pore radius. This was done for NRe = 2000 to

correspond with the experimental conditions under which the rejection was

measured by Geraldes et al. [94].

That the use of a non-spherical model yields a larger apparent pore radius than a

spherical pore model is a fairly obvious result, which is consistent with other

results in the literature [159, 160]. In addition, the ratio of the calculated pore

radius for the spherical model to the non-spherical model (~0.6) is similar to that

obtained by Kiso et al. [159, 160]. If the Hagen-Poiseuille equation is used to

predict the pure water flux through the membrane with an assumed porosity of 0.1

as per Kiso et al. [159], a pore length of 8.9 × 10-8 m yields the same

hydrodynamic permeability as measured by Geraldes et al. [94]. This is broadly

consistent with other stated pore lengths in the literature [e.g. 159].

189
The rejection factors calculated by the pore model in conjunction with the

hydrodynamic model are shown in Table 6.2. Rejection values shown are

averaged over the membrane surface for both inlet Reynolds numbers, using the

pore radius of 0.79 nm determined using the experimental data for the higher inlet

Reynolds number of NRe = 2000. Good agreement is shown between the

experimentally determined rejection factors and the averaged values obtained

through the present model. For PEG1000, the modelled molecular dimensions are

sufficiently large to ensure complete rejection along the entire axial length (that is,

the modelled molecular shape is larger than the entrance to the membrane pores

regardless of its orientation). The experimental rejection data indicates that

PEG1000 should have rejection slightly less than unity, which suggests that either

the modelled molecular shape is too large; or, more likely, that there is some

variability in pore size within the membrane so that a small proportion of solute is

permitted through the larger pores. This issue can be addressed with the use of a

pore size distribution rather than a uniform pore size to represent the porous

membrane structure as described by Kiso et al. [160], which could be included in

future work. The axial variation in rejection for both compounds is largely

negligible, as the variation in osmotic pressure and hydrodynamic pressure is

small compared to the applied TMP for the test cases.

190
Table 6.2 - Predicted and experimentally observed rejection factors for test compounds for CDNF50l thin-

film composite NF membrane (Separem, Biela, Italy)

7UDQVPHPEUDQHSUHVVXUHǻP
Compound 1 MPa 2 MPa 3 MPa 4 MPa
Robs* Rcal † Robs Rcal Robs Rcal Robs Rcal

0.993 0.993 0.993 0.993


Sucrose 0.993 0.996 0.998 0.999
0.993 0.993 0.993 0.993

1.000 1.000 1.000 1.000


PEG1000 0.9981 0.9987 0.9993 0.9995
1.000 1.000 1.000 1.000

* Observed rejection data from [94] for NRe = 2000



Calculated rejection data averaged over membrane surface using pore model derived from [159] with pore size of 0.79 nm
for NRe = 500 and NRe = 2000, respectively

The permeate fluxes obtained from these simulations are shown in Figure 6.2. The

modelled permeate fluxes are almost identical to those obtained using a constant

experimentally obtained rejection factor as in Figure 4.4. As well, the predicted

CP boundary layer profiles, shown in Figure 6.3, are largely identical to those

obtained with the experimental rejection factors shown in Figure 4.5.

191
Sucrose, NRe = 500 Sucrose, NRe = 2000
6 6

5 5

Permeate flux (×10 -5 m3 m-2 s-1 )


Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Transmembrane pressure (MPa) Transmembrane pressure (MPa)
PEG1000, NRe = 500 PEG1000, NRe = 2000

6 6

5 5
Permeate flux (×10 -5 m3 m-2 s-1 )

Permeate flux (×10 -5 m3 m-2 s-1 )

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Collector 1 model Collector 2 model Collector 3 model


Collector 1 experimental Collector 2 experimental Collector 3 experimental
Collector 1 Geraldes et al. Collector 2 Geraldes et al. Collector 3 Geraldes et al.

Figure 6.2 - Comparison of predicted permeate fluxes for CFD model with experimentally determined fluxes

of [94] using calculated spatially varying rejection factors from pore model

192
Sucrose PEG1000

0.05 0.05

NRe = 500

NRe = 500
0.04 0.04

Dimensionless CP layer height (įW / h)


Dimensionless CP layer height (įW / h)

0.03 NRe = 2000 0.03

NRe = 2000

0.02 0.02

0.01 0.01

-2.78E-17 -2.78E-17
0 10 20 30 0 10 20 30
Dimensionless axial length (x/h) Dimensionless axial length (x/h)

1 MPa model 1 MPa Geraldes et al.


2 MPa model 2 MPa Geraldes et al.
3 MPa model 3 MPa Geraldes et al.
4 MPa model 4 MPa Geraldes et al.

Figure 6.3 - Comparison of dimensionless CP boundary layer profile for CFD model with computational

results of [94] using calculated spatially varying rejection factors from pore model

6.2.3 Simulations for low transmembrane pressure

To see if the axial variation in rejection would have any effect on fluxes or CP

layer thicknesses using the experimental conditions of Geraldes et al. [94], a trial

was conducted only for a lower range of TMPs ( 0.125 – 0.5 MPa), using both an

axially varying rejection, and a constant rejection obtained from the average of the

prior simulation. This exercise has been performed for the sucrose solution only,

as the modelled pore dimensions ensure complete rejection of PEG1000

193
regardless of TMP. For this case, the axial variation in rejection is shown in

Figure 6.4 as a function of the non-dimensional axial length. The predicted

rejection decreases along the channel length, as the osmotic pressure increases and

the hydrodynamic pressure decreases. No results are shown for the higher

Reynolds number case for the TMPs of 0.125 MPa and 0.25 MPa, as the osmotic

pressure in these cases is large enough to preclude positive flux through the

membrane.

Sucrose, NRe = 500


1
Predicted rejection Rcal

0.99

0.98

0.97
0 20 40 60 80 100
Sucrose, NRe = 2000
1
Predicted rejection Rcal

0.99

0.98

0.97
0 20 40 60 80 100
Dimensionless axial length (x/h)

0.125 MPa 0.25 MPa 0.375 MPa 0.5 MPa

Figure 6.4 - Axial variation in sucrose rejection for low transmembrane pressure with spatially varying

rejection factor

194
Comparisons of the predicted membrane fluxes and CP layer profiles for the low

TMP cases are shown in Figure 6.5 and Figure 6.6, respectively. The predicted CP

layer profiles and fluxes are identical whether the rejection is varying or constant,

primarily due to the very small variations in rejection as shown in Figure 6.4.

However, if conditions were such that the axial rejection variation were greater, it

might be expected that the predicted CP layer thickness for the varying rejection

simulation is slightly decreased compared to the constant rejection. The

discrepancy in CP layer thickness would presumably grow larger as the axial

distance increases: as the rejection factor decreases along the channel length, the

hydrodynamic pressure decreases and osmotic pressure increases. Consequently,

as more solute is permitted through the membrane at increasing axial lengths, the

solute concentration at the membrane surface and hence the height of the CP

boundary layer would, in theory, decrease. Similarly, one might expect a slight

decrease in the modelled fluxes when the rejection is allowed to vary, which

becomes more significant as the axial distance along the membrane channel

increases, though this is not observed for the conditions modelled here.

195
Sucrose, NRe = 500 Sucrose, NRe = 2000
7 7

6 6

Permeate flux (×10 -6 m3 m-2 s-1 )


Permeate flux (×10 -6 m3 m-2 s-1 )

5 5

4 4

3 3

2 2

1 1

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.3 0.4 0.5 0.6
Transmembrane pressure (MPa) Transmembrane pressure (MPa)

Collector 1 varying Collector 2 varying Collector 3 varying


Collector 1 constant Collector 2 constant Collector 3 constant

Figure 6.5 - Comparison of predicted permeate fluxes for CFD model for low transmembrane pressures using

calculated constant and spatially varying rejection factors

Sucrose
0.08

0.07
NRe = 500

0.06
Dimensionless CP layer height (įW / h)

0.05

0.04

0.03

NRe = 2000
0.02

0.01

0
0 20 40 60 80 100
Dimensionless axial length (x/h)

0.125 MPa varying 0.125 MPa constant


0.25 MPa varying 0.25 MPa constant
0.375 MPa varying 0.375 MPa constant
0.5 MPa varying 0.5 MPa constant

Figure 6.6 - Comparison of dimensionless CP boundary layer profile for CFD model for low transmembrane

pressures using calculated constant and spatially varying rejection factors

196
The inability of the spatially variable rejection factor to address the discrepancy

between modelled and experimentally measured fluxes suggests that further

mechanisms not included in the current model may be operating to reduce the

flux, separate from any variation in the intrinsic rejection of the membrane.

Determining the nature of these mechanisms and quantifying how they may be

represented without resorting to the empirical approach of the osmotic pressure

correction factor may be a useful direction of investigation for future work.

6.3 Coupled modelling of solutions containing charged solutes

To incorporate the rejection model described in Chapter 5 for charged solutes, a

significant amount of additional numerical calculations are required within the

hydrodynamic model. To do so within the native Command Expression Language

(CEL) of the ANSYS CFX software is impractical, hence the capabilities of user

Fortran are required. This is a facility of the CFX software where external code

can be written (in Fortran 77 and / or Fortran 90) and compiled to interface with

the CFX solver [123]. This allows for a great deal of extensibility to the

hydrodynamic model, though the actual coding process is to some extent tedious,

not entirely straightforward and requires significant investment of time. This

section will briefly describe the development of the user Fortran interface

developed to couple the hydrodynamic model and the rejection model.

197
6.3.1 Model details

The simplest method available to interface the CFX solver with external code is

the use of so-called User Routines. In this method, CEL expressions within the

CFX model are defined in terms of User Functions, which call User Routines

consisting of libraries compiled from the external Fortran code. The arguments

that are required to be passed to the external routine are described in Table 6.3.

Table 6.3 – Minimum arguments to be passed from CFX solver to external Fortran routine for calculating

rejection of charged solutes

Arguments Comments

Parameters specific to each solute

Mass fraction Determined at the membrane surface by the hydrodynamic model

Molecular weight User-supplied property

Valence User-supplied property

User-supplied property that can be over-written if molecular length


Stokes radius
and width are supplied

User-supplied property that can be over-written if molecular length


Diffusivity
and width are supplied

Molecular length User supplied property; if not supplied spherical pore model used

Molecular width User supplied property; if not supplied spherical pore model used

Parameters specific to the membrane

Membrane pore radius User-supplied property

Membrane charge valence User-supplied property

Membrane charge density User-supplied property

Hydrodynamic parameters

Volumetric flux Determined at the membrane surface by the hydrodynamic model

198
This method has some limitations on its generality as follows:

x The User Functions defined within the CFX model must have a fixed

number of scalar arguments. This means, for example, that separate User

Functions (and effectively separate CFX models) must be developed for

systems with two uncharged solutes, three uncharged solutes and so on.

x The User Routine is allowed to return only one scalar argument, which in

this instance is the solute rejection for one of the charged solutes. For a

system of only two charged solutes, this is feasible as the rejection of both

solutes must be identical to maintain electroneutrality, and so the second

solute rejection can simply be set to that of the first solute. However, for a

system of n > 3 solutes, the rejections are in general not identical, and so

to return the n unique rejections, n separate User Functions must be

constructed and the numerical routine must be run n times. This is clearly

not ideal as it makes the model construction convoluted, and more

importantly adds significantly to the computational expense.

These limitations could be overcome if the User Routines were allowed to accept

and return array-valued (vector) arguments, as per the Matlab sample code shown

in Appendix A. However, at this stage this functionality is unfortunately not

possible in ANSYS CFX. Nevertheless, the User Routine approach can be made

to work satisfactorily, examples of which are presented in the following.

199
The general programming approach is similar to the MATLAB model presented

in Appendix A, with separate Fortran subroutines (written in Fortran 90) used to

separately determine steric partitioning coefficients, Donnan equilibrium

coefficients, and solve the BVP numerically. To interface with the CFX solver, a

‘wrapper’ subroutine must be written in Fortran 77 to pass the input arguments to

the Fortran 90 subroutines and finally return the calculated rejection to the CFX

solver. Sample code for calculating the rejection of a sodium chloride solution is

shown in Appendix B, and a flowchart of the computational process is shown in

Figure 6.7. The computational procedure is broadly similar to that for the

uncharged case, except that the external Fortran routine is called to calculate

rejection, and additional algebraic variables must be created to record the

calculated rejections and permeate fluxes within the CFX results file, as results

returned from User Routines are not automatically incorporated as such.

200
Figure 6.7 – Flowchart of computation procedure for coupled hydrodynamic-mass transfer model for charged

solutes only

201
6.3.2 Special considerations for the coupled model

Several points should be made regarding the successful coupling of the

hydrodynamic model and the rejection model. These issues largely concern the

passing of the solute mass fractions (obtained from the hydrodynamic model

solute including concentration polarisation effects) to the external rejection

routine. Perhaps most important is the physical locale in the computational

domain where the rejection calculations are actually performed. In principle, the

coupled approach is capable of determining the rejection separately for every

single computational cell adjacent to the membrane surface, depending on the

local hydrodynamic conditions. Previous simulations in Section 6.2.2 indicated

that when permitted, solute rejection did not vary appreciably over the membrane

surface for the relatively high transmembrane pressures common in nanofiltration

and reverse osmosis applications. In addition, solving the BVP defined by the

Nernst-Planck equations to perform the rejection calculations separately at each

computational point on the membrane surface is more expensive. Despite these

drawbacks, for these coupled simulations, the rejection has been allowed to vary

spatially as described in the interests of obtaining the most general and rigorous

solution possible. If desired, however (perhaps for reasons of computational

expense or when negligible variation in rejection can be expected), an averaged or

area-averaged solute mass fraction may be passed to the external rejection routine,

so that the rejection is calculated as a constant value across the entire membrane

surface.

202
Another point of interest in coupling the models lies in the iterative solution

scheme used by CFX when solving a multicomponent fluid problem. During the

first ‘false time-step’, the solute mass fractions at the membrane surface may be

uniformly zero if the solute mass fractions are not explicitly specified throughout

the domain in the initial conditions (i.e. the membrane channel is effectively

initially filled with clean water). This causes numerical difficulties for the coupled

rejection routine. In some cases it may be desirable to specify a ‘clean-water’

initial condition, so to avoid the numerical difficulties this causes, the rejection

routine is simply bypassed until an appreciable (non-zero) solute mass fraction

appears at the membrane surface within the hydrodynamic model.

6.3.3 Model verification – two charged solutes

Using the code presented in Appendix B, the coupled model was run to reproduce

the results of Geraldes et al. [94], previously modelled using an user-assigned

rejection in Chapter 4. To do so, it was first necessary to fit appropriate membrane

parameters to the experimental rejection data so that the membrane parameters in

the pore model could be assigned.

Real rejection data for the CDNF50l membrane used by Geraldes et al. were given

for the two charged test compounds of sodium chloride and sodium sulphate [94]

as functions of the applied transmembrane pressure. The rejection values for

sodium sulphate are all close to unity, while those for sodium chloride vary

between about 0.4 and 0.7. Therefore, the data for sodium chloride were used to

203
fit the membrane parameters, as the near-unity rejections for sodium sulphate

mean that the model is relatively insensitive to variation in the parameters.

An additional difficulty in fitting the membrane parameters for this case is that no

experimental flux data was given by Geraldes et al. [94] for the sodium chloride

experiments. Instead, the applied transmembrane pressures are listed, and it is

stated that solute wall concentrations from their CFD model, using the

experimentally determined permeate fluxes as a computational boundary

condition, are used to predict the real rejections. In this case, neither the

experimental permeate fluxes for sodium chloride or the wall concentrations

determined by Geraldes et al. [94] are known. To work around this, computational

results from the CFD model presented in Chapter 4, averaged over the entire

membrane area, have been used to estimate the permeate fluxes and wall

concentrations, as shown in Table 6.4

Table 6.4 – Estimates for permeate fluxes and wall concentrations for NaCl solution based on previous CFD

results

Transmembrane pressure
Parameter
1 MPa 2 MPa 3 MPa 4 MPa
-1 -3 -3 -3
NaCl mass fraction (kg kg ) 2.39×10 3.28×10 4.33×10 5.70×10-3
NaCl concentration (mol m-
3 40.8 56.0 73.9 97.3
)
Permeate flux (m3 m-2 s-1) 1.14×10-5 2.54×10-5 3.94×10-5 5.34×10-5

Using the values shown in Table 6.4, the non-linear least squares fitting procedure

described in Chapter 5 was again used to fit membrane parameters to the rejection

204
data for sodium chloride. The obtained parameters are shown in Table 6.5, and a

comparison between the model and experimental data is shown in Figure 6.8.

Some comments are necessary regarding the fit, as the adopted parameters are

unusual: firstly, the fitted value for the pore radius is very small (0.29 nm), and is

inconsistent with the fitted value previously obtained for the same membrane

using uncharged solutes (~0.79 nm). Secondly, the membrane charge density is

very low compared to previous fitting results for other membranes. A good fit to

the experimental data was not possible using larger pore radius or membrane

charge density values. The very low membrane charge density is perhaps most

easily explainable: it may simply be that this particular NF membrane does not

exhibit significant charge effects.

However, the small pore radius is more problematic; it would be expected to be

similar to the pore radius to that previously obtained using uncharged solutes. One

explanation for the discrepancy is that the estimates used for the wall

concentration and permeate flux, obtained from previous CFD model results, do

not correspond well with those used by Geraldes et al. [94]. However, as the

values used by Geraldes et al. are unknown, these values represent a reasonable

‘best guess’. A simple sensitivity analysis was conducted on the wall

concentration and permeate flux values by increasing and decreasing the value of

each by ± 100 %, as shown in Table 6.5. The analysis indicated that this large

variation in the concentration and flux values had no effect on the fitted pore

radius, and a relatively small effect on the thickness parameter ǻy/Ak. However, the

205
fitted charge density CX was somewhat sensitive to the variation in wall

concentration. From this it is reasonable to conclude that the fitted value for the

pore radius of 0.29 nm, while very small, is acceptable, and that the previous

fitting for uncharged solutes is less applicable, as the rejections for the uncharged

solutes are generally near unity and hence less sensitive to the fitting process.

Table 6.5 – Fitted membrane parameters for CDNF50l membrane for NaCl solution

Parameters ǻy/Ak (×10-6 m) rp (nm) CX (×10-3 eq m-3)

Fitted values 4.34 0.29 1.3

Fitted values, flux +100% 1.93 0.29 0.95

Fitted values, flux -100% 7.05 0.29 0.75

Fitted values, wall 4.34 0.29 54.4


concentration+100%

Fitted values, wall concentration- 4.34 0.29 1.1


100%

206
100

90
Numerical - with steric effects
Experimental
80

70

60
Ri (%)

50

40

30

20

10

0
1 2 3 4 5 6
5 3 -2 -1
Flux Jv × 10 (m m s )

Figure 6.8 – Comparison of predicted rejection and experimental rejection for NaCl solution with CDNF50l

membrane using fitted parameters from Table 6.5

The coupled model was then run using these fitted parameters for the lower

entrance Reynolds number of 500 for transmembrane pressures of 1, 2, 3, and 4

MPa as per previous simulations.

The results of the simulations are shown in Figure 6.9 in terms of the predicted

rejection and permeate flux when averaged over the membrane area. The spatial

variation in the predicted rejection is also shown in Figure 6.10. A small decrease

in rejection is apparent along the axial flow direction, which becomes less

significant as the transmembrane pressure is increased. This rejection variation

207
along the channel is in the order of 5 % for an applied transmembrane pressure of

1 MPa, decreasing to approximately half a percent for the 4 MPa case. The

variation is less apparent at higher transmembrane pressures as the retarding effect

of the osmotic pressure developed within the concentration polarisation boundary

layer is less significant compared to the applied pressure.

The rejections predicted by the coupled model correspond almost exactly to the

previous predictions by the pore model only, which confirms the external

rejection routine is producing reasonable results. The comparison to the

experimental rejection data is essentially the same as before, and produces a

reasonable fit.

It is interesting to note however that the permeate fluxes predicted by the coupled

model for the given transmembrane pressures are slightly less than the permeate

fluxes extrapolated from previous (uncoupled) CFD model results, as shown in

Table 6.4. The reason for this disparity is due to the previous CFD model results

assuming a constant rejection for the entire membrane channel, whereas the

current (coupled) model allows for the rejection to vary spatially across the

membrane surface. However, refitting the pore model with the modified fluxes

predicted by the coupled model produces almost no variation in the fitted

membrane parameters, as would have been expected based on the sensitivity

analysis previously conducted (shown in Table 6.5).

208
100

90

80

70

60
Ri (%)

50

40

30 Numerical - with steric effects


Experimental

20 CFD model results

10

0
1 2 3 4 5 6
Flux Jv × 105 (m3 m-2 s -1)

Figure 6.9 – Comparison of experimental rejection (unfilled markers), predicted rejection from pore model

(solid line) and predicted rejection from coupled CFD-pore model (filled markers) for NaCl solution with

CDNF50l membrane using fitted parameters from Table 6.5

209
1 MPa
0.490
Predicted rejection Rcal

0.480

0.470

0.460

0.450

0.440
0 10 20 30 40 50 60 70 80 90 100

2 MPa
0.615
Predicted rejection Rcal

0.610
0.605
0.600
0.595
0.590
0.585
0 10 20 30 40 50 60 70 80 90 100

3 MPa
0.660
Predicted rejection Rcal

0.655

0.650

0.645

0.640
0 10 20 30 40 50 60 70 80 90 100

4 MPa
0.678
Predicted rejection Rcal

0.676

0.674

0.672

0.670

0.668
0 10 20 30 40 50 60 70 80 90 100
Dimensionless axial length (x/h)

Figure 6.10 – Predicted axial rejection variation from coupled CFD-pore model for NaCl solution with

CDNF50l membrane using fitted parameters from Table 6.5

The results can also be analysed once again in terms of the observed concentration

polarisation boundary layers as per Section 4.2.5. A comparison of the

concentration polarisation layer thicknesses predicted by the present coupled

model and the model in [94] is shown in Figure 6.11. Despite the discrepancy

210
between the predicted and experimental rejections, the concentration polarisation

layer thicknesses remain close to that predicted by Geraldes et al. [94] – i.e. the

polarisation layer is less sensitive to the rejection than it is to the hydrodynamic

conditions within the membrane channel. Nevertheless, the coupled approach is

surely a step forward in a rigorous method for simulating concentration

polarisation and mass transfer, and this comparison confirms that the coupled

model approach still allows for the concentration polarisation boundary layer to

be developed accurately.

0.06

0.05
Dimensionless CP layer height (įW / h)

0.04

0.03
1 MPa model
1 MPa Geraldes et al.
2 MPa model
0.02 2 MPa Geraldes et al.
3 MPa model
3 MPa Geraldes et al.
4 MPa model
4 MPa Geraldes et al.
0.01

0
0 10 20 30
Dimensionless axial length (x/h)

Figure 6.11 - Comparison of dimensionless CP boundary layer profile for coupled CFD-pore model with

computational results of [94]

211
6.3.4 Simulations – multiple charged solutes

The coupled model is also capable of predicting the hydrodynamic and rejection

behaviour of systems of an arbitrary number of charged solutes. However,

detailed hydrodynamic measurements of such multicomponent systems (such as

the concentration polarisation boundary layer formation and the variation in solute

concentrations through the membrane) are not available in the literature. In

addition the costs of specialised apparatus to measure and visualise such systems

accurately are prohibitively expensive for this project. Thus, a series of numerical

experiments have been performed for solutions with multiple charged solutes for

several reasons:

x To verify the rejection calculations produced by the coupled model for

multiple charged solutes agree with previous calculations and

experimental data; and

x To investigate general trends and hydrodynamic phenomena that arise

from the coupled hydrodynamic / multiple solute approach, which have

not been previously investigated.

We first attempt to verify the coupled hydrodynamic rejection model against the

rejection results of Bowen and Mukhtar [17] for mixtures of sodium chloride and

sodium sulphate solutions using the PES5 membrane. These were previously

modelled using the rejection model only in Section 5.5.2.2.

212
6.3.4.1 Verification (without concentration polarisation)

To use the coupled model, the hydrodynamics of the membrane channel must be

described. The experimental setup used in [17] was that of a cylindrical stirred

cell (an Amicon Model 8200 cell of 200 mL capacity) – the stirring of the cell

minimises the formation of the concentration polarisation layer. To explicitly

model the hydrodynamics of the cell caused by the motion of the stirrer would be

prohibitively difficult, so it is assumed that the cell can be well described by a 2D

membrane channel with a permeable lower wall, with negligible concentration

polarisation occurring. The lack of concentration polarisation is modelled by

simply setting the transport properties of the solutes (the ions) to be identical to

that of the solvent (water) for the bulk hydrodynamic simulation.

The channel has equivalent height to the cell (approximately 66 mm), and a

channel length equal to that of the cell diameter (62 mm). The fluid velocity

across the membrane surface within the stirred cell may be assumed to be small,

and so can be modelled by setting a small inlet velocity for the fictive 2D channel.

For these simulations, the inlet Reynolds number was set to an arbitrary small

value of 10. It is stressed that the intent of these simulations is not to construct an

exhaustive hydrodynamic model of the stirred-cell arrangement. Rather, the intent

is to provide broadly similar hydrodynamic conditions by which the external

rejection routine may be tested by comparing the average rejections across the

membrane surface. For these purposes, the simplified geometric and flow

configuration described here is sufficient.

213
In addition, the hydrodynamic permeability of the membrane is required. The

permeability has been obtained from pure water flux experiments conducted by

Bowen and Mukhtar [17]; our analysis of the data indicates the permeability of

the PES5 membrane to be approximately 7.76 × 10-11 m².s kg-1.

The model was then run with the parameters previously described in Section

5.5.2.2 for a range of transmembrane pressures as used in [17]. A detailed listing

of the simulation parameters is shown in Table 6.6.

Table 6.6 – Model input parameters for simulations corresponding with Bowen & Mukhtar [17]

Parameter Value

Membrane channel length L 62 mm

Membrane channel half height h 33 mm

Membrane channel width w Variable depending on mesh density – equal to size of


smallest mesh element

Transmembrane pressure 50, 100, 200, 300, 400 kPa

Membrane permeability LV 7.76 x 10-11 m Pa-1 s-1

Inlet Reynolds number NRe 10

Membrane pore radius rp 1.14 nm

Solution bulk concentration CB 1 mol m-3

Molar ratio (NaCl:Na2SO4) 0.2, 0.4, 0.6, 0.8

Membrane charge density CX 6.80, 6.39, 5.96, 5.50 mol m-3 *

0HPEUDQHWKLFNQHVVSDUDPHWHUǻy/Ak 1.60 × 10-5 m

* Variable according to solution ionic strength; described by isotherm in Eqn. (5.30) with parameters
s = 0.523, q = 0.699.

214
6.3.4.2 Verification results (without concentration polarisation)

An initial verification of the coupled model results can be obtained by comparing

the rejections measured and predicted in [17] to the rejections predicted by the

coupled model, when averaged over the membrane surface. This comparison is

shown in Figure 6.12, with the area-averaged rejections from the coupled

hydrodynamic / rejection model shown with filled markers. The results indicate

good agreement with the experimental data, indicating that the simplifying

assumptions made about the cell hydrodynamics are justifiable, and that the

coupled rejection routine returns reasonable results for multiple charged solutes.

There is a small discrepancy between the flux values predicted by the coupled

model and the experimentally measured values for each transmembrane pressure.

The values of the transmembrane pressures used for the experimental

measurements in [17] are not explicitly stated, but it may be reasonably assumed

that the pressures were as listed in Table 6.6 (50, 100, 200, 300, and 400 kPa),

since they correspond with the values listed in [17] used to determine the

hydrodynamic permeability of the membrane. In any case, this discrepancy in flux

is not significant in this case, as the intent is simply to verify the validity of the

rejection routine within a reasonable amount of accuracy. It is likely that the flux

discrepancy simply reflects the simplifying assumptions made about the

hydrodynamics of the stirred cell. Further, it is reasonable to infer that if a

rigorous attempt to model the stirred cell hydrodynamics were made, it would

215
result in higher velocities across the membrane surface, thus increasing the

predicted permeate flux, which would reduce the apparent discrepancy.

100 100
(a) (b)
80 80
60 60
40 40
20 20
Ri (%)

Ri (%)
0 0
-20 -20
-40 -40
-60 -60
-80 -80
0 10 20 30 40 0 10 20 30 40
6 3 -2 -1 6 3 -2 -1
Flux Jv × 10 (m m s ) Flux Jv × 10 (m m s )

100 100
(c) (d)
80 80
60 60
40 40
20 20
Ri (%)

Ri (%)

Na + model
Cl- model
0 0
SO2- model
4

-20 -20
+
Na experimental
-
Cl experimental
SO2-
-40 -40 4
experimental

Na + CFD
-60 -60 Cl- CFD
SO2-
4
CFD

-80 -80
0 10 20 30 40 0 10 20 30 40
6 3 -2 -1 6 3 -2 -1
Flux Jv × 10 (m m s ) Flux Jv × 10 (m m s )

Figure 6.12 - Comparison of experimental rejection data from Bowen and Mukhtar [17], numerically

computed rejections from pore model, and area-averaged predicted rejections from coupled hydrodynamic /

pore model for PES5 membrane for 1×10-3 M salt solution with molar ratios of: (a) 0.2, (b) 0.4, (c) 0.6, (d)

0.8 (NaCl:Na2SO4)

216
Even though concentration polarisation effects are neglected for this model, some

spatial variation in rejection over the membrane surface is apparent for the

relatively low transmembrane pressures encountered here. The variation in

rejection for these cases is generally in the order of several percent. This is due to

the axial decrease in flux along the membrane channel due to pressure drop from

the permeable wall condition (as discussed in Section 4.2.3). Some representative

graphs of this rejection variation are shown for a TMP of 200 kPa, with a molar

ratios (NaCl:Na2SO4) of 0.2 in Figure 6.13. The variation shown here is typical of

the other cases simulated.

Additionally, the pressure drop through the membrane channel should cause an

axial decrease in flux along the membrane surface. This turns out to be almost

negligible for these conditions in the absence of concentration polarisation – the

variation is typically less than half a percent of the average flux. A representative

example of the spatial variation in permeate flux is shown in Figure 6.14 for the

molar ratio of 0.2, which is typical of the other cases simulated.

217
Na+
0.800

0.795
Predicted rejection Rcal

0.790

0.785

0.780

0.775
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
SO42-
0.900

0.895
Predicted rejection Rcal

0.890

0.885

0.880

0.875

0.870
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Cl-
-0.200

-0.220
Predicted rejection Rcal

-0.240

-0.260

-0.280

-0.300

-0.320

-0.340
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.13 – Predicted axial rejection variation for PES5 membrane for 1×10-3 M salt solution with molar

ratios of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa

218
25

20 50 kPa
100 kPa
200 kPa
300 kPa
Predicted permeate flux J V × 10 6 (m3 m-2 s-1 )

15

10

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.14 – Predicted axial permeate flux variation for PES5 membrane for 1×10-3 M salt solution with

molar ratio of 0.2 (NaCl:Na2SO4)

The results of the verification study in the absence of concentration polarisation

indicate the coupled rejection model is working as expected The non-negligible

axial variation in solute rejection indicates that use of the spatially variable

coupled rejection model is desirable in order to perform a rigorous simulation of

219
the membrane system, particularly at lower transmembrane pressures. The

coupled model may now be extended fully by considering the effects of

concentration polarisation by re-including spatially variable, concentration-

dependent transport properties as described in the following section.

6.3.4.3 Simulations with concentration polarisation effects

We now perform a series of numerical experiments for the coupled multiple

charged solute model with concentration polarisation effects. First, we revisit the

simple hydrodynamic model of the stirred cell used in [17] as described in Section

6.3.4.2, but now add a description of the mixed solution transport properties to

include the phenomenon of concentration polarisation. This allows us to compare

the relative effect of concentration polarisation for a known case. A more

extensive set of simulations for more realistic membrane channel geometries will

be performed in Chapter 7. This allows for identification of general features of the

hydrodynamics and mass transfer predicted by the coupled model for a typical

membrane channel.

The use of mixed salt solutions provides some additional difficulty in estimating

the transport parameters of the mixed solution. As discussed in Section 3.4,

rigorous simulation of the diffusion of components in solution using the Maxwell-

Stefan approach [125] or the Onsager approach [126] is possible, but largely

infeasible due to the lack of experimental data for determining diffusion

coefficient matrices. In addition, the use of the Nernst-Planck equations to

220
describe the diffusion of components within the bulk solution has been shown to

cause severe numerical difficulties when coupled with a commercial CFD model

[103]. In particular, the only way such a multi-component diffusion model can be

incorporated for the entire flow domain in ANSYS CFX is through use of an

additional source term as used by Fimbres-Weihs and Wiley [103], which was

shown to cause numerical difficulties. For these reasons, a simplified Fick’s law

approach has been adopted for these simulations, and is detailed in the following.

Previously relationships established by Geraldes et al. [94] were used to describe

the viscosity, osmotic pressure, and diffusivity of sodium chloride and sodium

sulphate solutions as functions of the salt mass fractions. It is difficult to find

references in the literature that describe similar relationships for mixed solutions

of sodium chloride and sodium sulphate. For these simulations some empirical

approximations have been made to estimate these relationships as follows.

The first property to consider is the diffusivity of the solution. In principle, the

diffusivity of each ion in the system could be described independently as per the

Fick’s law approximation; however, preliminary tests indicated that this made it

difficult to ensure the electroneutrality condition within the bulk solution and at

the membrane surface was met. Instead, the multicomponent fluid was considered

to consist of four separate ions (Ion 1: Na+, Ion 2: Cl-, Ion 3: Na+, Ion 4: SO42).

The diffusivity of Ions 1 and 2 was given by the relationship derived by Geraldes

et al. [94] for sodium chloride, and that of Ions 3 and 4 was given by the similar

relationship for sodium sulphate. That is:

221
DIons 1,2 u109 (m 2s -1 ) 1.61>1 - 14 ZIon 1  ZIon 2 @ for ZIon 1  ZIon 2 < 0.006;
(6.1)
1.45 for ZIon 1  ZIon 2 > 0.006

>
DIons 3,4 u109 (m 2s -1 ) 1.23 1 - 0.76 ZIon 3  ZIon 4
0.4
@ (6.2)

Effectively this represents the scenario where the salts do not dissociate in

solution, and each salt diffuses independently of the other. While this is of course

not a rigorous representation of the real ionic transport within the bulk flow, it is a

reasonable approximation to make in light of the difficulties encountered in a full

multi-component diffusion model for the bulk flow. This ensures that the

electroneutrality condition is upheld, which allows the predictive rejection model

described in Section 5.3.3 to determine the mass transfer through the membrane in

terms of a source applied at the boundary only (rather than throughout the entire

domain). If possible, further work to determine the variation in mixed salt solution

diffusivity as a function of concentration would be beneficial.

To describe the solution viscosity, an average is made of the viscosities calculated

using the relationships described by Geraldes et al. [94]:

0.89>1 - 1.63 ZIon 1  ZIon 2 @  0.89>1 - 3.52 ZIon 3  ZIon 4 @


Psolution u103 (Pa s) (6.3)
2

Again, this is satisfactory for this investigation, though further description of the

viscosity variation of multicomponent solutions would be of future interest.

222
Finally, the osmotic pressure exerted by the mixed salt solution is simply taken as

the sum of the osmotic pressures exerted by the sodium chloride and sodium

sulphate solutions in isolation:

3 solution u 10-5 (Pa ) 805.1 ZIon 1  ZIon 2  337.8 ZIon 3  ZIon 4


0.95
(6.4)

The extension to more complex multicomponent solutions is fairly obvious,

provided relationships are available for the variation in these transport properties

as functions of the mass fractions of each solute in isolation. Clearly this is not as

rigorous a description of the transport of ions within the bulk membrane flow as

would be provided by a Nernst-Planck approach as per Fimbres-Weihs and Wiley

[103], but on the other hand it avoids the large computational expense in doing so,

and the associated numerical difficulties as the governing equations become very

stiff. A flowchart of the computational procedure when using this simplified

diffusion model for multicomponent flow is shown in Figure 6.15.

223
Figure 6.15 – Flowchart of computation procedure for coupled hydrodynamic-mass transfer model for

multiple charged solutes with simplified diffusion model

224
Besides the approach for describing the solution transport properties shown

above, all parameters (e.g. geometries, boundary conditions etc) used for the

simulations with concentration polarisation are identical to those in Section

6.3.4.1. For brevity the simulation results presented here are only for Case (a),

with a molar ratio (NaCl:Na2SO4) of 0.2.

6.3.4.4 Simulation results (with concentration polarisation)

The coupled model including concentration polarisation effects was generally able

to produce well-converged solutions. However, the increased stiffness of the mass

transport equations due to the inclusion of the concentration polarisation effects

leads to a fairly significant increase in computational expense: simulations with

identical meshes and computational parameters generally required at least three

times (and sometimes up to 20 times) as many iterations to converge to a steady-

state solution, compared to the case with no concentration polarisation.

However, it was difficult to obtain a converged solution for the highest

transmembrane pressure simulated of 400 kPa, even with very refined meshes and

small (false) timestepping. Internal model computations appeared to indicate for

this transmembrane pressure that the rejection of the chloride co-ion varied axially

between relatively large negative and positive rejection values (approximately

-0.28 to 0.25). This appears to have caused severe numerical difficulties for the

CFX solver in using the rejection to calculate the mass source / sink term to be

applied at the membrane boundary. That is, with the large axial variation in

225
rejection causing both positive and negative flux of chloride ions along the same

boundary in the hydrodynamic model, the mass transfer equations appear to have

become very stiff. It should be noted that this particular scenario (where large

variations in both positive and negative rejection are observed for the same ion

within a membrane channel depending on the axial position) is in part reflective

of the low crossflow velocity adopted for these simulations, and is probably

unlikely to occur in practical membrane channels. However, further work to

determine the cause of the numerical failure of the solver routine would be

beneficial.

The computational results indicate that the effect of the concentration polarisation

boundary layer is very significant at the low crossflow velocity investigated here.

A comparison of the predicted area-averaged rejections from the coupled CFD

model both with and without concentration polarisation effects is shown in Figure

6.16. While the predicted fluxes are slightly less for a given transmembrane

pressure, there is a significant reduction in rejection when concentration

polarisation effects are included. This is due to the increased viscosity and

osmotic pressure near the membrane surface, which retard the permeation through

the membrane; and due to the increased solute mass fractions at the membrane

surface, which reduce the real rejection of the membrane.

(It should be noted that the results from the coupled CFD model with

concentration polarisation are not intended to be compared to the experimental

results in this case. The large discrepancy between the predicted rejections using

226
the coupled CFD model with concentration polarisation, and the experimental

results, is not a reflection of any failing on the part of the CFD model. This is

because the CFD model has not attempted to model the complex hydrodynamics

of the stirred cell arrangement as previously discussed, which in the absence of

concentration polarisation is an acceptable approximation. Clearly, when CP

effects are included, this approximation is invalid. Instead, this comparison is

intended to illuminate the effect of including concentration polarisation on

rejection behaviour by comparing a given (theoretical) case without concentration

polarisation, and the same (theoretical) case with concentration polarisation.)

227
100
(a)
(d)
80

60

40

20
Ri (%)

Na + model
0 Cl- model
SO24 - model
Na + experimental
-20 Cl- experimental
SO24 - experimental
Na + CFD
-40 Cl- CFD
SO24 - CFD

Na + CFD with CP
-60 Cl- CFD with CP
SO24 - CFD with CP

-80
0 5 10 15 20 25 30 35 40
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 6.16 - Comparison of experimental rejection data from Bowen and Mukhtar [17], numerically

computed rejections from pore model, and area-averaged predicted rejections from coupled hydrodynamic /

pore model both with and without concentration polarisation effects for PES5 membrane for 1×10-3 M salt

solution with molar ratio of 0.2 (NaCl:Na2SO4)

It is also instructive to examine the axial variation in the predicted rejection and

permeate flux with concentration polarisation included in the CFD model.

Previously, in the absence of concentration polarisation, the axial rejection

variation was in the order of several percent. With concentration polarisation

accounted for, the axial variability in rejection is much more pronounced at the

channel inlet as the concentration polarisation layer begins to develop. In some

228
cases, the variation is as much as 25 to 30 percent for the positively rejected ions

(Na+ and SO42-). Some representative graphs of this rejection variation are shown

for a TMP of 200 kPa, in Figure 6.17.

The predicted axial variation in permeate flux for varying transmembrane

pressures is shown in Figure 6.18. Compared to the results without concentration

polarisation, the decrease in flux along the axial length of the channel is much

more pronounced. The variation is as great as approximately 10% of the average

permeate flux for the lowest simulated transmembrane pressure (50 kPa). This

indicates that the inclusion of concentration polarisation effects can have

significant effects on the predicted permeate flux, as well as the predicted

rejection, and can further explain the large variability in predicted rejection in the

case. (That is, due to the axial decrease in permeate flux, the predicted rejections

also decrease axially in accordance).

229
Na+
0.850

0.800
Predicted rejection Rcal

0.750

0.700

0.650

0.600

0.550

0.500
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
SO42-
0.950
0.900
Predicted rejection Rcal

0.850
0.800
0.750
0.700
0.650
0.600
0.550
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Cl-
-0.200
-0.250
Predicted rejection Rcal

-0.300
-0.350
-0.400
With CP
-0.450
Without CP
-0.500
-0.550
-0.600
-0.650
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.17 – Predicted axial rejection variation for PES5 membrane both with and without concentration

polarisation effects for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa

230
25

300 kPa

20
Predicted permeate flux J V × 10 6 (m3 m-2 s-1 )

200 kPa
15

With CP
Without CP
10

100 kPa

50 kPa

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.18 – Predicted axial permeate flux variation for PES5 membrane both with and without

concentration polarisation effects for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4)

231
It is also interesting to examine the computed concentration fields for each ion,

including the concentration polarisation boundary layer, in the presence of a

negatively rejected ion. A selection of representative graphs of the variation in

concentration in the vicinity of the membrane surface (normalised relative to the

bulk inlet concentration) are shown in Figure 6.19. For the positively rejected

ions, the classical concentration polarisation boundary layer is formed above the

membrane surface as expected. However, for the negatively rejected chloride co-

ion, an ‘inverse’ concentration polarisation boundary layer is formed, where the

concentration at the membrane surface is less than that in the bulk of the cell (the

spectrum between blue and purple representing the inverse concentration

polarisation layer).

232
(a)

(b)

(c)

Figure 6.19 – Detail of computed concentration fields (normalised relative to bulk inlet concentration) of (a)

Na+, (b) Cl-, (c) SO42- for PES5 membrane with concentration polarisation effects for 1×10-3 M salt solution

with molar ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa

233
This is shown in more detail in Figure 6.20, which shows the variation in solute

concentration (normalised relative to the inlet bulk concentration) in the normal

direction to the membrane surface, at the centre of the cell. The concentration of

the negatively rejected chloride co-ion is seen to increase with increasing distance

from the membrane surface. Figure 6.20 also illustrates that the thickness of the

concentration polarisation layer also varies slightly between ions due to their

varying transport properties. This suggests that, while not as rigorous as a full

ionic diffusion model for the bulk flow (used with partial success by [103]), the

coupled CFD-mass transfer model using the diffusivities of individual salts may

offer a more physical representation of the concentration polarisation layer than a

simple empirical correlation (e.g. a film layer model).

1.700
Normalised mass fraction (Ȧi /Ȧi,B)

1.500

Cl-
Na+
1.300
SO42-

1.100

0.900

0.700
0.001 0.01 0.1
Dimensionless normal distance from membrane (y/2h)

Figure 6.20 – Predicted variation normal to membrane surface for solute concentration normalised relative to

inlet bulk concentration for PES5 membrane with concentration polarisation effects for 1×10-3 M salt solution

with molar ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa

234
6.4 Coupled modelling of mixed solutions

We have seen how the coupled hydrodynamic / rejection model approach is able

to describe in detail the behaviour of membrane systems with arbitrary numbers of

charged or uncharged solutes. For a practical feed solution, there may be a

combination of charged and uncharged solutes; in this section, the previous

modelling approaches are combined with further User Fortran to provide a general

predictive model which addresses this situation.

6.4.1 Model details

The implementation of the combined charged-uncharged solute model is fairly

straightforward. Again, to ensure electroneutrality conditions are met and

avoiding the numerical effort and difficulty of a full diffusion model for the entire

flow domain, a simplified Fick’s law approach is used to describe diffusion of

salts as described in Section 6.3.4.3. Similarly for uncharged solutes, the diffusion

of each uncharged solute is assumed to occur independently of other solutes

(charged or uncharged), and hence the diffusivity coefficient in water for each

uncharged solute is used to describe this. The independent diffusion of an

arbitrary number of uncharged solutes therefore does not violate the

electroneutrality requirement for the charged solute rejection routine. Little data is

available to describe the diffusion of multiple uncharged solutes in a single

solution, so this approach is a reasonable approximation. It is noted however that

in principle the diffusion of charged and uncharged solute could also be described

235
by a rigorous diffusion model for the entire flow domain, as in [103], though the

computational difficulties in this approach remain prohibitive. In addition, this is a

potential area for future research which may yield some benefit for membrane

CFD modelling.

The approach used to describe the viscosity of the solution and the osmotic

pressure generated is the same as that of Section 6.3.4.3. That is, expressions for

the viscosity and osmotic pressure as a function of the mass fraction for single-

solute solutions (corresponding to the components of the multicomponent

solution) are used, and:

x The calculated viscosities are averaged to obtain a representative viscosity

for the multicomponent solution; and

x The exerted osmotic pressures are summed to obtain an overall osmotic

pressure.

A flowchart demonstrating the computational procedure for this general model

with a mixed solution of charged and uncharged solutes is shown in Figure 6.21.

236
Figure 6.21 – Flowchart of computation procedure for coupled hydrodynamic-mass transfer model for

multiple charged and/or uncharged solutes with simplified diffusion model

237
To demonstrate the model, the scenario used in Section 6.3.4.3 has been revisited

(the simplified stirred-cell arrangement of Bowen and Mukhtar [17] including

concentration polarisation effects for a NaCl:Na2SO4 mixture). However, an

additional uncharged solute (sucrose) has been included to demonstrate the

efficacy of the numerical routine and to briefly examine the predicted effects on

the membrane channel behaviour. The properties used for the additional

uncharged solute (sucrose) are shown in Table 6.7.

Table 6.7 – Properties for additional uncharged solute (sucrose) in coupled simulation of charged-uncharged

solute mixture

Parameter Value

Solute length LS 1.137 nm

Solute width WS 0.513 nm

Bulk mass fraction + 0.0001 kg kg-1 (~ 0.291 mol m-3)

Single-solute viscosity * 0.89(1 + 1.31Ȧi + 16.83Ȧ i 2) × 103 (Pa s)

Single-solute diffusivity * 0.52(1 í 1.33Ȧi) × 109 (m2 sí1)

Single-solute osmotic pressure * 72.18Ȧi(1 + 0.94Ȧi + 2.93Ȧ i 2) × 10í5 (Pa)

* Relationship obtained from [94].


+
Value selected to be similar to concentration of salts.

6.4.2 Simulations

Simulations were run for the conditions previously described in Section 6.3.4.3

with the addition of sucrose as an uncharged solute. For brevity, the results

presented here are only for the molar ratio (NaCl:Na2SO4) of 0.2. Well converged

solutions were obtained for all cases, even for the highest transmembrane pressure

of 400 kPa, which previously proved difficult to obtain converged solutions for in

238
the case examined in Section 6.3.4.3. The reason for the improved numerical

stability with the addition of an uncharged solute is unknown and difficult to

speculate upon.

A comparison of the area-averaged predicted rejections and permeate fluxes for

the various transmembrane pressures (50 kPa, 100 kPa, 200 kPa, 300 kPa, 400

kPa) is shown in Figure 6.22. Here a comparison is made to the case presented in

Section 6.3.4.4 to examine the additional effect of the uncharged sucrose molecule

on the rejection of the other (charged) components and the volumetric permeate

flux. Examination of Figure 6.22 reveals that the addition of sucrose has a

retarding effect on the permeate flux in all cases. This is an expected result, as the

higher viscosity within the concentration polarisation layer from the addition of

sucrose, as well as the additional osmotic pressure exerted by the sucrose

molecules, both act to reduce the permeate flux in accordance with Eq. (4.1).

239
100

80

60

40
Sucrose
Na +
20 Cl
-
Ri (%)

SO2 -
4

Na + without sucrose
0 -
Cl without sucrose
SO2 - without sucrose
4

-20

-40

-60

-80
0 5 10 15 20 25 30 35 40
6 3 -2 -1
Flux Jv × 10 (m m s )

Figure 6.22 - Comparison of area-averaged predicted rejections from coupled hydrodynamic / pore model

with concentration polarisation effects for PES5 membrane for 1×10-3 M salt solution with molar ratio of 0.2

(NaCl:Na2SO4) with and without sucrose addition

The variation in the predicted rejection between the two cases is also

commensurate with this trend. If the rejection model is considered in isolation, the

calculated ionic rejections should remain constant with the addition of any

number of uncharged solutes. However, when considered in the coupled

hydrodynamic-mass transfer model, the differing hydrodynamic conditions within

the membrane channel and particularly within the concentration polarisation layer

present a different set of inlet boundary conditions to the ionic rejection model.

240
That is, the inclusion of the uncharged solutes within the hydrodynamic model

yields slightly different velocity, pressure, and concentration fields, which affect

the coupled rejection calculations. This is expressed in the simulation results by

the magnitudes of the predicted rejections in the presence of sucrose being slightly

less than predicted for a purely electrolytic solution. Specifically, the rejections of

the positively rejected ions (Na+ and SO42-) are smaller positive values when

sucrose is added, and the rejections of the negatively rejected Cl- co-ion are

smaller negative values when sucrose is added. Finally, as expected, the predicted

rejection of sucrose monotonically increases with increasing transmembrane

pressure.

An examination of the axial variation in the predicted rejections is shown in

Figure 6.23, again only for the transmembrane pressure of 200 kPa. The general

trend in axial rejection variation with sucrose addition is very similar to that

without sucrose, though the addition of sucrose tends to slightly reduce the

rejection as previously noted. There is also some axial variation in the rejection of

sucrose itself, though this variation is small in magnitude compared to that of the

charged solutes.

A comparison of the axial variation in permeate flux is also shown in Figure 6.24.

Again, the rapid initial decrease in flux near the channel in inlet is observed,

followed by a very gradual decrease along the remainder of the channel. As

previously noted, the addition of sucrose causes a slight decrease in flux due to

241
the higher viscosity of the mixed solution and the greater osmotic pressure

exerted.

A representative detail of the concentration fields predicted for each solute, again

normalised relative to the inlet concentration, in the vicinity of the membrane

surface is shown in Figure 6.25. Again, the inverse concentration polarisation

layer for the negatively rejected chloride co-ion is observed, while the relative

variation in concentration is observed to be smaller for sucrose than the other

components. The variation in solute concentration (normalised relative to the inlet

concentration) in the normal direction to the membrane surface, at the centre of

the cell is also shown in Figure 6.26. This shows the variation in the thickness of

the concentration polarisation for sucrose compared to the other components more

clearly. The variation for the other components remains similar to that observed in

the absence of sucrose.

242
Na+
0.730
0.710
Predicted rejection Rcal
0.690
0.670
0.650
0.630
0.610
0.590
0.570
0.550
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
SO42-
0.800

0.750
Predicted rejection Rcal

0.700

0.650

0.600

0.550
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
-
Cl
-0.550
-0.560
Predicted rejection Rcal

-0.570 With sucrose


-0.580 Without sucrose
-0.590
-0.600
-0.610
-0.620
-0.630
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Sucrose
0.708

0.707
Predicted rejection Rcal

0.706

0.705

0.704

0.703
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.23 – Predicted axial rejection variation for PES5 membrane for 1×10-3 M salt solution with molar

ratio of 0.2 (NaCl:Na2SO4) for TMP of 200 kPa with and without sucrose addition

243
35
With sucrose
Without sucrose

30 400 kPa

25
Predicted permeate flux J V × 10 6 (m3 m-2 s-1 )

300 kPa

20

15 200 kPa

10

100 kPa

5
50 kPa

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Dimensionless axial length (x/h)

Figure 6.24 – Predicted axial permeate flux variation for PES5 membrane both with and without sucrose

addition for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4)

244
(a)

(b)

(c)

(d)

Figure 6.25 – Detail of computed concentration fields (normalised relative to inlet bulk concentration) of (a)

Na+, (b) Cl-, (c) SO42-, (d) sucrose for PES5 membrane for 1×10-3 M salt solution with molar ratio of 0.2

(NaCl:Na2SO4) with sucrose addition for TMP of 200 kPa

245
1.700
Normalised mass fraction (Ȧi /ȦiB)

1.500

Cl-
Na+
1.300
SO42-
Sucrose
1.100

0.900

0.700
0.001 0.01 0.1
Dimensionless normal distance from membrane (y/2h)

Figure 6.26 – Predicted variation normal to membrane surface for solute concentration normalised relative to

inlet bulk concentration for PES5 membrane for 1×10-3 M salt solution with molar ratio of 0.2 (NaCl:Na2SO4)

with sucrose addition for TMP of 200 kPa

6.5 Conclusions

A method for coupling the CFD simulations of membrane channel hydrodynamics

(discussed in Chapter 4) and the mass-transfer model for prediction of solute

rejection (discussed in Chapter 5) has been successfully developed and verified.

The method is capable of providing a reasonably general and rigorous model of

the high-pressure membrane filtration processes (NF and RO) that is fully able to

predict the salient behaviours of such a system; i.e. both the channel

hydrodynamics, which significantly influence concentration polarisation and

fouling behaviour and hence permeate flux, and the mass transfer of solutes

through the membrane, which affects permeate quality. However, these processes

246
do not act in isolation, and interact in complex and subtle ways. The ability of the

coupled model to capture these interactions makes it a powerful predictive tool for

modelling of membrane filtration.

Simulations for several simple cases were carried out to verify the predictions of

the coupled model, which revealed some interesting (predicted) phenomena. The

model allows for both the permeate flux and the rejection of individual solutes to

vary spatially over the computational domain, according to the local

hydrodynamic conditions. Axial reduction in permeate flux through a membrane

channel has been widely described and observed, but axial reduction in solute

rejection less so. Model results indicated that axial rejection variation was so

small to be negligible in some cases, while quite significant in other cases at very

low crossflow velocities. While these velocities were generally considerably

smaller than encountered in practical applications, the utility of including spatially

variable solute rejection is apparent.

The use of the coupled rejection model in conditions where negative ionic

rejection was observed also allowed prediction of an inverse concentration

polarisation layer (i.e. where concentrations at the membrane surface are less than

that of the bulk flow). While this is a fairly obvious consequence of negative

rejection, this has not been widely described in CFD models of membrane

filtration.

247
The coupled model also accounts in part for the interplay in rejection and

permeate flux between uncharged and charged solutes. If a DSPM-style rejection

model is considered in isolation, the rejections of charged and uncharged

molecules are essentially independent. However, by coupling the hydrodynamic

model predictions of the concentration polarisation layer formation to the

rejection routine, the addition of uncharged solutes to a salt solution reduces both

the overall permeate flux and rejection of the charged solutes due to the increased

viscosity of the concentration polarisation layer, and the increased osmotic

pressure exerted within. This would also similarly affect other uncharged solutes.

In this manner, this gives a more physical representation of the complex

concentration polarisation layer mass transfer behaviour than has typically been

presented in CFD simulations of the membrane filtration process.

Of course, several simplifying assumptions have been made in the development of

the model, either for reasons of numerical stability or mathematical convenience.

These represent natural areas of further progression for the modelling approach

which would be valuable to investigate in further studies. The first of these

assumptions (discussed previously in Chapter 5) is the neglect of dielectric

exclusion effects in the modified DSPM used to predict solute rejection. As

highlighted earlier, inclusion of these effects would provide a more general

rejection model (at the expense of some further mathematical and modelling

difficulty), though the numerical method used in the coupled model is well-suited

to extension to include these effects.

248
The method to describe the diffusion of solutes throughout the bulk flow in the

membrane channel is also relatively simplified, using a modified Fick’s law

approach to ensure electroneutrality is conserved throughout the domain. The

reasoning for this approach is as follows: while rigorous diffusion models for

mixed solutions exist, they suffer from either a lack of experimental data to

determine diffusion coefficient matrices (e.g. an Onsager [126] or Maxwell-Stefan

[125] type approach), or numerical difficulties when implemented via a fictive

source term in commercial CFD codes (e.g. the extended Nernst-Planck approach

attempted in [103]). The combination of mean salt diffusion coefficients for

charged solutes and individual diffusion coefficients for uncharged solutes

ensures electroneutrality conditions are conserved, and incurs comparatively small

computational expense and numerical difficulty. The downfall of this approach is

it relies on empirical combination of experimentally determined relations

describing the transport properties of single-component solutions, and is as such

not completely general. However, model results indicated that this approach was

still able to simulate the formation of component-specific concentration

polarisation layers, which is thought to provide a more physical representation of

the concentration polarisation phenomena than simple film theory style

approximations. Further work in both predicting transport and rheological

properties of mixed solutions, and implementing a rigorous electro-diffusive

model for the bulk channel flow, would be important steps towards a more general

model.

249
Finally, some numerical instabilities were observed for an isolated case using the

coupled model that mesh and timestep refinement were unable to resolve. The

reason for this is unknown and is difficult to speculate upon, but appears to be

related to the spatially variable rejection routine predicting both positive and

negative flux for a single ion over a single boundary source. This condition is

unlikely to occur in realistic operating conditions, but further investigation and

refinement of the numerical routine if necessary would be worthwhile.

We may now turn our attention to using the coupled hydrodynamic-mass transfer

model to conduct some more detailed simulations of high-pressure membrane

filtration channels in Chapter 7.

250
Chapter 7

Simulations using the coupled hydrodynamic and rejection

model

7.1 Introduction

In Chapter 6, a fully predictive method coupling a hydrodynamic (CFD) model

and a rejection (pore) model was described, with verifications and some

illustrative simulations carried out for simple hydrodynamic cases. These

verifications were only possible for such simple cases due to the inherent

difficulties in observing and measuring both the flow structures for practical

membrane geometries, and the concentration fields within the membrane channel

with complex feed solutions. Verifications were thus performed for these simple

cases using larger-scale spatially averaged parameters (such as the rejection and

permeate flux), which indicated the coupled model performed well.

The coupled model is also able to predict the behaviour of the membrane system

in a level of detail that is (for most purposes) not presently feasible by

experimental observation. In this chapter the coupled model is used to conduct

some detailed numerical experiments for more realistic membrane channel

geometries and hydrodynamic conditions. The purpose of these simulations is to

identify some of the salient features of the hydrodynamics and mass transfer

251
predicted by the coupled model for a typical membrane channel. In the absence of

detailed experimental data at these fine scales for such systems, such simulations

may provide a useful tool for predicting the behaviour and performance of

membrane systems.

7.2 Simulation approach

All simulations conducted in this chapter are for a 2D model of a spacer-filled

membrane channel. Previously in Section 4.3, a hydrodynamic model of such a

channel was developed using the geometry and flow conditions previously

modelled by Wardeh and Morvan [115]. These conditions provided for several

complex hydrodynamic flow features to be observed, and are thus an interesting

hydrodynamic case to investigate further. In the following simulations, we use the

previous hydrodynamic model geometry and flow conditions, but extend the

model by inclusion of the coupled mass transfer model to predict the rejection of

solutes by the membrane.

In the previous simulations of the spacer-filled channel, a simple sodium chloride

solution was considered with a high fixed rejection (R = 0.95) specified at the

membrane boundary. It is noted that this value was used as an arbitrary value in

numerical simulations of a (fictive) membrane only. To provide an insight into the

behaviour of the coupled model, we now consider a case where the mass-transfer

behaviour of the system is more variable; i.e. where the solute rejection varies

over a wider range of values. To do so, we adopt membrane parameters similar to

252
that used in Chapter 5 for the PES5 membrane used in experiments by Bowen and

Mukhtar [17].

Finally, we perform the simulations for several different feed solutions, including

single salt solutions, multiple salt solutions, and multicomponent mixtures

containing both salts and organic (uncharged) solutes. In this way, we aim to

provide a general investigation of the utility of the coupled model, identify

relevant hydrodynamic and mass transfer phenomena, and evaluate areas for

future investigation beyond this thesis. A brief revision of the simulation

parameters is contained in the following.

7.2.1 Geometry

As per Section 4.3.3, the model geometry used by Wardeh and Morvan [115] is

employed, consisting of a 2D channel with long entrance and exit lengths. The

channel is filled with seven cylindrical spacers oriented normal to the main

channel flow, which are arranged in a zig-zag fashion (alternating between the top

and bottom of the channel) to emulate a woven spacer net. Both upper and lower

walls of the channel are permeable membranes, with the exceptions of the

surfaces of the spacer filaments and the entrance and exit sections. The geometry

is used to represent the periodic flow that occurs in spiral-wound membrane

(SWM) modules; the use of seven filaments is sufficient to develop the periodic

flow structure in the vicinity of the fifth and six filaments, as previously noted. A

253
recap of the model geometry is shown in Figure 7.1, with the dimensions h = 2

mm, l/h = 4, and df / h = 0.5.

spacer filaments
channel height h
diameter df permeate flow

feed retentate

entrance length permeate flow exit length


12 df 25 df

inter-filament distance l

Figure 7.1 – Model geometry for simulations of flow through 2D space-filled channel

7.2.2 Boundary and initial conditions

The same boundary and initial conditions as described in Section 4.3.4 are

enforced here, with the average velocity (in terms of the channel Reynolds

number Rech) and solute mass fraction(s) specified at the channel inlet, and an

average static zero pressure specified at the outlet. No-slip walls with no mass

transfer are used over the entrance and exit lengths, and on the filament surfaces.

The permeate flow through the membrane surfaces is again specified as a mass

sink, though the solute fluxes are now determined by the coupled mass transfer

model described in Chapter 6, which calculates the rejection of each individual

solute independently for each computational point on the membrane surface. The

properties of the membrane are derived from the parameters used in Chapter 5 for

the PES5 membrane used in experiments by Bowen and Mukhtar [17].

254
Finally, only the two lower channel Reynolds numbers previously considered

(Rech = 100 and 200) are simulated here, for reasons of computational expense.

This is because higher channel Reynolds numbers (Rech •   ZHUH VKRZQ LQ

Section 4.3.7.2 to cause unsteady flow structures, greatly increasing

computational expense, even with an uncoupled (hydrodynamics only) model.

Due to the much greater computational expense again of the coupled

(hydrodynamics and mass transfer) modelling approach, a decision was made to

investigate steady flows only for the coupled model, though future investigation

of unsteady flows would indeed be worthwhile. Initial conditions are again

automatically generated by the CFX solver, with the initial solute concentration

field set to be globally equal to that of the bulk solute mass fraction. The model

parameters common to all subsequent simulations are thus shown in Table 7.1.

Table 7.1 – Model input parameters for coupled simulations of 2D spacer-filled channel

Parameter Value

Transmembrane pressure 0.5 MPa

Membrane permeability LV 2 x 10-10 m Pa-1 s-1

ǻy/Ak 1.60 ×10-5 m

CX 15 eq m-3 (fixed ionic strength for all simulations)

rp 1.14 nm

ZX -1 (negatively charged membrane)

7.2.3 Computational mesh and convergence criteria

An unstructured 2D mesh was used for all simulations, with significant refinement

of the mesh in the vicinity of the channel walls and the surfaces of the spacer

255
filaments. Previous mesh refinement studies for the same geometry without the

coupled mass transfer model indicated a spatial resolution of approximately 0.05

mm in the bulk of the channel was required, decreasing down to approximately 3

μm against the membrane walls and filaments. However, with the inclusion of the

coupled mass transfer model, a slightly more refined mesh was required to obtain

accurate solutions. The adopted mesh consists of 0.02 mm elements in the channel

bulk, with refinement down to 2 μm against the membrane walls and filaments.

The refined mesh consists of approximately 358 000 elements (cf. 295 000

elements for the mesh previously adopted in Section 4.3.6).

Convergence was judged to occur once the mass, momentum, and solute mass

fraction residuals were all smaller than 1 x 10-5. This convergence target is relaxed

compared to previous simulations as it proved difficult to obtain convergence to

the very tight previous tolerance of 1 x 10-6 with the coupled model.

7.3 Coupled simulation of single salt solution in spacer filled channel

The first set of simulations was conducted for a single salt solution of sodium

chloride using the parameters outlined in Table 7.1 and Figure 7.1. An inlet salt

mass fraction of 0.002 kg kg-1 was used for these simulations, representing an

ionic strength of approximately 34.1 eq m-3 (i.e. an inlet sodium chloride

concentration of 34.1 mol m-3). Simulations were performed for channel Reynolds

numbers of 100 and 200; at these Reynolds numbers, the flow observed was

steady, and there was no variation of the solution with time.

256
A selection of representative results from the simulations are shown here, again

all in the vicinity of the fifth and sixth spacer filaments (chosen as a representative

monitoring location). As before, the predicted salt mass fraction (Figure 7.2), fluid

streamlines (Figure 7.3), velocity vectors (Figure 7.4), and vorticity fields (Figure

7.5) are shown. The salt mass fraction shown in Figure 7.2 has been normalised

relative to the inlet (bulk) salt mass fraction for comparison with later cases.

Rech = 100

Rech = 200

Figure 7.2 – Predicted salt mass fraction distribution for spacer-filled channel using coupled-mass transfer

model for single salt solution (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

257
Rech = 100

Rech = 200

Figure 7.3 – Predicted fluid streamlines for spacer-filled channel using coupled-mass transfer model for

single salt solution (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

Rech = 100

Rech = 200

Figure 7.4 – Predicted velocity vectors for spacer-filled channel using coupled-mass transfer model for single

salt solution (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

258
Rech = 100

Rech = 200

Figure 7.5 – Predicted vorticity fields for spacer-filled channel using coupled-mass transfer model for single

salt solution (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

The general features of the flow are similar to that previously observed in Section

4.3.7.1. Large recirculation zones are formed on the downstream face of the

filaments, and smaller zones on the upstream faces, in which correspondingly

high salt concentrations are observed due to concentration polarisation. These

zones grow larger with increasing channel Reynolds number. However, as the salt

rejection for the current case (R varying between approximately 0.017 – 0.018) is

significantly less than the case in Section 4.3.7.1 (R = 0.95), the magnitude of the

variation in salt concentration throughout the channel is greatly reduced. The

concentration polarisation effect is still observable, however.

259
Of more interest is the variation in mass transfer parameters within the monitoring

area due to the inclusion of spatially variable rejection and permeate flux. The

predicted variation in the salt rejection, permeate flux, and wall shear are shown in

Figure 7.6 and Figure 7.7, for Rech = 100 and Rech = 200, respectively. The

predicted salt mass fraction and velocity streamlines are shown for spatial

reference. Values are shown both at the upper and lower (permeable) surfaces of

the membrane channel.

260
Salt mass fraction

Velocity streamlines

Rejection R
0.01784
Lower wall

0.01776 Upper wall

0.01768

0.01760

Flux JV (× 10-5 m³ m² s-1)


6.65
Lower wall

6.55 Upper wall

6.45

6.35

Wall shear (Pa)


0.40
Lower wall
0.30 Upper wall

0.20
0.10
0.00

Figure 7.6 - Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3, Rech = 100)

261
Salt mass fraction

Velocity streamlines

Rejection R
0.01750
Lower wall

0.01730 Upper wall

0.01710

0.01690

Flux JV (× 10-5 m³ m² s-1)


6.30
Lower wall
6.20 Upper wall

6.10
6.00
5.90

Wall shear (Pa)


1.00
Lower wall
0.80
Upper wall
0.60
0.40
0.20
0.00

Figure 7.7 - Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3, Rech = 200)

262
Considering first the predicted wall shear, the maximum shear is observed as

expected opposite to the spacer filaments. That is, the maximum shear for the

lower wall is observed at the same axial location as the filament on the opposing

(upper) wall, and vice versa. This result corresponds with previous observations

for similar channels [e.g. 114]. A local increase in wall shear is also observed

adjacent to the recirculation zones on the downstream side of the filaments, due to

the recirculation motion previously described (clockwise rotation adjacent to the

lower wall, and counter-clockwise adjacent to the upper wall). A similar local

increase in shear might be expected adjacent to the upstream filament edge due to

the smaller recirculation zone observed here; however, for the relatively low

channel Reynolds numbers examined here, the increase in shear in this region is

not significant.

The axial variation in permeate flux, however, follows a different trend from that

of the wall shear. Flux decreases on the downstream side of an adjacent filament,

as the recirculation zone separates from the membrane wall However, flux then

increases towards the upstream side of an adjacent filament, as the recirculation

zone reattaches to the membrane wall. The region of highest flux (i.e. the highest

mass transfer) is located slightly downstream of the reattachment point, which is

consistent with the results of Fimbres-Weihs et al. [114]. The flux variation

implies, as noted in [114], that separation points and reattachment points are

associated with regions of low and high mass transfer, respectively.

263
Perhaps somewhat surprisingly, the spatial variation in salt rejection follows an

almost identical trend to that of the permeate flux. Examining the rejection

variation in conjunction with the predicted salt mass fraction distribution, this can

be succinctly explained by the reasoning presented in [114]; at reattachment

points, low-concentration fluid flows towards the (high-concentration) wall.

Simultaneously, flow within the boundary layer adjacent to the wall moves away

from the reattachment point, creating a low salt concentration and thereby

increasing the predicted rejection. Similarly, a high salt concentration is formed at

separation points, decreasing the predicted rejection. However, the variation in

rejection is relatively small in comparison to the variation in permeate flux.

The formation of the concentration polarisation boundary layer can also be

visualised more clearly by examining the variation in salt mass fraction

(normalised relative to the inlet concentration) in the direction normal to the

membrane wall. This has been performed for a point midway between filaments

in the monitoring zone in terms of non-dimensionalised distance from the

membrane surface for both the upper and lower membrane surfaces. The variation

is shown in Figure 7.8 for the lower membrane surface, and Figure 7.9 for the

upper membrane surface. Clearly in the single salt case this transverse variation in

concentration will be identical for both ions; however it is useful for examining

the formation of varying concentration polarisation layers for different ions in a

multicomponent solution, as will be discussed in Section 7.4.

264
1.014
Rech = 100
1.012 Rech = 200
Normalised mass fraction (Ȧi /ȦiB)

1.010

1.008

1.006

1.004

1.002

1.000
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.8 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from lower membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

1.014
Rech = 100
1.012 Rech = 200
Normalised mass fraction (Ȧi /ȦiB)

1.010

1.008

1.006

1.004

1.002

1.000
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.9 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from upper membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl, 0.002 kg kg-1, equivalent to 34.1 mol m-3)

265
As expected, Figure 7.8 and Figure 7.9 show the magnitude of the transverse

variation in concentration to be greatest at the lower channel Reynolds number.

That is, the effect of concentration polarisation is greatest for the lower Reynolds

number. Additionally, the concentration polarisation effect is more pronounced

for the upper membrane wall (i.e. on the wall opposite to the immediately

downstream filament). This reflects the disparity in permeate flux between the

lower and upper walls at the filament midpoint shown in Figure 7.6 and Figure

7.7.

Finally, to examine the effect of allowing salt rejection to vary spatially using the

coupled model, an additional set of simulations was conducted using the same

parameters, but with a manually specified constant rejection (i.e. a non-coupled

model). In these cases, the rejection was specified to be equal to the rejection

determined by the coupled model simulations, averaged over the monitoring area

in the vicinity of the fifth and sixth filaments. That is, for the Rech = 100 case, the

rejection was set to a constant value of 0.0178, and for the Rech = 200 case, a

value of 0.0172.

Examination of the results of the constant rejection (non-coupled) simulations

revealed the results to be practically identical to that of the coupled model, and

therefore the results are not reproduced here. This is due to the relatively low

spatial variation in rejection for these conditions.

266
7.4 Coupled simulation of multiple salt solution in spacer filled

channel

A second set of simulations was then conducted for a multiple salt solution

composed of sodium chloride and sodium sulphate with an arbitrarily chosen

molar ratio of 0.2:0.8 (NaCl:Na2SO4). The inlet mass fraction was adjusted to

maintain a consistent ionic strength with the previous simulation for a single salt

solution in Section 7.3 (approximately 34.1 eq m-3); this yields inlet

concentrations of approximately 3.79 mol m-3 (NaCl) and 15.16 mol m-3

(Na2SO4). (Maintaining a consistent ionic strength between simulations allows for

the use of a consistent membrane volumetric charge density CX for easier

comparison between cases).

Simulations were again run for two channel Reynolds numbers of 100 and 200,

and again steady flow was observed in all cases. Of particular interest is the fact

that negative rejection of the chloride co-ion is predicted for these conditions,

similar to the conditions presented in Section 5.5.2.2. Ionic mass fraction

distributions are shown, normalised relative to the inlet mass fraction for each ion,

in Figure 7.10 and Figure 7.11 for Rech = 100 and Rech = 200 respectively. These

clearly indicate the negative rejection of the chloride co-ion and the formation of

the ‘inverse’ concentration polarisation layer – the colour spectrum between pink

and blue on these figures represents areas where the local concentration is less

than the inlet concentration (i.e. where the normalised mass fraction is less than

267
one). As before, for the positively rejected ions (Na+ and SO42-), the classical

concentration polarisation layer is formed near the membrane surface, and areas

of high concentration are formed adjacent to the spacer filaments due to the

recirculation zones in these regions. Again, the high concentration zones are more

prominent on the downstream face of the filaments due to the relatively larger size

of the corresponding recirculation zone. Corresponding low concentration regions

for the negatively rejected Cl- ion occur in these recirculation zones.

Na+

Cl-

SO42-

Figure 7.10 – Predicted ionic mass fraction distribution for spacer-filled channel using coupled-mass transfer

model for multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3),

Rech = 100

268
Na+

Cl-

SO42-

Figure 7.11 – Predicted ionic mass fraction distribution for spacer-filled channel using coupled-mass transfer

model for multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3),

Rech = 200

Predicted streamlines, velocity vectors, and vorticity plots are again shown in

Figure 7.12, Figure 7.13, and Figure 7.14, respectively. These show the same

general hydrodynamic trends as observed for the single salt case discussed in

Section 7.3.

269
Rech = 100

Rech = 200

Figure 7.12 – Predicted fluid streamlines for spacer-filled channel using coupled-mass transfer model for

multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3)

Rech = 100

Rech = 200

Figure 7.13 – Predicted velocity vectors for spacer-filled channel using coupled-mass transfer model for

multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3)

270
Rech = 100

Rech = 200

Figure 7.14 – Predicted vorticity fields for spacer-filled channel using coupled-mass transfer model for

multiple salt solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3)

We can again examine the spatial variation in mass transport in terms of solute

rejection, permeate flux, and wall shear, as well. These are shown in Figure 7.15

and Figure 7.16 for Rech = 100 and Rech = 200, respectively, with fluid

streamlines shown for spatial reference. Examining first the solute rejections, a

significant variation is predicted depending on axial position between the

filaments. For the positively rejected ions (Na+ and SO42-), these follow a similar

trend to that observed for the rejection for the single salt solution discussed in

Section 7.3. Rejection decreases on the downstream side of an adjacent filament,

increases on the upstream side of an adjacent filament, and reaches a minimum

slightly upstream of an adjacent filament. Somewhat differently to the single salt

results, however, the rejection for these ions is at a maximum adjacent to an

271
opposing filament, rather than the previously observed location slightly

downstream of the reattachment point for the downstream filament recirculation

zone. This is more apparent at the higher channel Reynolds number as shown in

Figure 7.16. A local maximum is however still observed at the reattachment point

for the downstream filament recirculation zone for the lower channel Reynolds

number, as shown in Figure 7.15.

For the negatively rejected co-ion (Cl-), the opposite of this trend is observed, as

might be expected. Maximum rejection occurs at the upstream filament

recirculation zone, decreasing rapidly towards the filament face, then increases

rapidly at the downstream face of the filament. The rejection then decreases

gradually, reaching a minimum adjacent to the opposing filament, again more

prominently at the higher channel Reynolds number. A local minimum

corresponding to the reattachment point of the downstream filament recirculation

zone is also observable for the Rech = 100 case.

Inspection of the predicted wall shear variation reveals very similar results to that

for the single salt solution, both in terms of general variation and shear magnitude.

This is perhaps expected as the bulk hydrodynamic conditions are in general very

similar.

More interestingly, and perhaps initially counter-intuituively, is the variation in

permeate flux. While this follows the same general trend as described for the

single salt solution in Section 7.3 (similar to that of the rejection trend for

272
positively rejected ions), the overall magnitude of the permeate flux is greater for

the multiple salt solution, as shown in Table 7.2. This represents a flux increase of

approximately 6 – 7 %, which is not insignificant.

Table 7.2 – Comparison of average permeate flux values for single salt (NaCl, overall ionic strength 34.1 eq

m-3) and multiple salt (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3) cases

Solution Averaged permeate flux JV × 10-5 m3 m-2 s-1

Rech = 100 Rech = 200

Single salt (NaCl) 6.51 6.16

Multiple salt (NaCl:Na2SO4) 6.92 6.58

This occurs despite identical operating parameters (inlet velocity, applied TMP)

and ionic strength being imposed. This result seems even more unintuitive

considering the considerable increase in the magnitude of rejections (in the order

of 0.2 – 0.25) for the multiple salt case, compared to the single salt case (in the

order of 0.02) – it might be expected that the higher rejection would lead to a

more highly developed concentration polarisation boundary layer, retarding the

permeate flux due to viscous effects and an increase in osmotic pressure.

However, the flux increase can be explained by several factors: first, the constant

ionic strength (expressed in eq m-3) used between simulations and the inclusion of

the multivalent sulphate ion means that the inlet salt concentration for the multiple

salt case (expressed in mol m-3) is less than the single salt case. This implies that

the osmotic pressure developed in the multiple salt case would be reduced, and

273
hence have a lesser retarding effect on the permeate flux. In addition, the overall

solution viscosity would be reduced to have a similar effect. More significant

perhaps, is the negative rejection of the chloride ion, which causes the formation

of the ‘inverse’ concentration polarisation layer, where the concentration of

chloride decreases in increasing proximity to the membrane surface. This reduces

solution viscosity and osmotic pressure near the membrane surface, which may

also lead to the increase in overall permeate flux. In any case, the permeate flux

increase illustrates the fact that subtle changes in the mass transfer behaviour of

the near-wall region may have significant effects on the flux behaviour of the

membrane which are not immediately or intuitively obvious without use of a

rigorous simulation tool.

274
Velocity streamlines

Na+ rejection
0.210
Lower wall
0.205 Upper wall

0.200
0.195
0.190

Cl- rejection
-0.180
Lower wall
-0.190 Upper wall

-0.200
-0.210
-0.220

SO42- rejection
0.250
Lower wall
0.240 Upper wall

0.230
0.220
0.210

Flux JV (× 10-5 m³ m² s-1)


7.10
Lower wall
7.00 Upper wall

6.90
6.80
6.70

Wall shear (Pa)


0.40
Lower wall
0.30 Upper wall

0.20
0.10
0.00

Figure 7.15 - Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1

eq m-3, Rech = 100)

275
Velocity streamlines

Na+ rejection
0.210
Lower wall

0.200 Upper wall

0.190

0.180

Cl- rejection
-0.170
Lower wall
-0.180
Upper wall
-0.190
-0.200
-0.210
-0.220

SO42- rejection
0.250
Lower wall
0.240 Upper wall

0.230
0.220
0.210

Flux JV (× 10-5 m³ m² s-1)


7.00
Lower wall
6.80
Upper wall
6.60
6.40
6.20
6.00

Wall shear (Pa)


1.00
Lower wall

Upper wall

0.50

0.00

Figure 7.16 - Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1

eq m-3, Rech = 200)

276
Examining the transverse variation in solute concentration can also give a clearer

picture of the concentration polarisation predictions of the model. Figure 7.17

through Figure 7.20 illustrate the variation in the normalised ionic concentrations

with the non-dimensionalised transverse distance. As previously discussed, this

analysis has been performed for both the upper and lower channel walls at a point

midway between the fifth and sixth spacer filaments. As discussed in Section

6.3.4.4, these results can be loosely interpreted as describing the overlapping

concentration polarisation layers of different thickness for each individual ion, or

more precisely, the varying concentration gradients for each individual ion within

the concentration polarisation layer. In all cases, the greatest variation in

concentration is observed for the SO42- ion. The inverse concentration polarisation

layer for the negatively rejected Cl- co-ion is also apparent. The magnitude of

variation in the non-dimensionalised concentration is also greater for the lower

channel Reynolds number, i.e. the concentration polarisation effect is stronger at

lower Reynolds number, as expected.

In addition, the variation is more pronounced (the concentration polarisation

effect is stronger) for the upper membrane wall. That is, the concentration

polarisation effect midway between filaments is stronger on the opposing wall

from the downstream filament than the adjacent wall. This is commensurate with

the permeate flux predictions shown in Figure 7.15 and Figure 7.16, and the

previous results for the single salt solution.

277
1.200

1.150 Cl-

Na+
Normalised mass fraction (Ȧi /ȦiB)

1.100
SO42-

1.050

1.000

0.950

0.900

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.17 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from lower membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 100)

1.200

1.150 Cl-

Na+
Normalised mass fraction (Ȧi /ȦiB)

1.100
SO42-

1.050

1.000

0.950

0.900

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.18 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from lower membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 200)

278
1.250

1.200 Cl-

Na+
Normalised mass fraction (Ȧi /ȦiB)

1.150
SO42-
1.100

1.050

1.000

0.950

0.900

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.19 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from upper membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 100)

1.250

1.200 Cl-

Na+
Normalised mass fraction (Ȧi /ȦiB)

1.150
SO42-
1.100

1.050

1.000

0.950

0.900

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.20 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from upper membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 200)

279
Again, to examine the effect of the spatial rejection variation inherent in the

coupled model approach, a set of simulations was conducted using fixed constant

rejections derived from the averaged predicted rejections from the coupled model.

A summary of the constant rejection values used in these simulations is shown in

Table 7.3.

Table 7.3 – Fixed rejection values used for uncoupled model comparison case (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3)

Component Rejection

Rech = 100 Rech = 200

Na+ 0.200 0.200

Cl- -0.205 -0.201

SO42- 0.233 0.232

Previous results for the single salt solution in Section 7.3 had indicated negligible

difference between the coupled simulation results with variable rejection, and the

uncoupled results with constant rejection. This was ascribed to the low variation

in rejection, and low overall rejection, predicted for the conditions in Section 7.3.

In the multiple salt case considered here, though, an observable difference in some

results is apparent. This is due to the higher rejection magnitudes predicted for the

present case, and hence the greater effect exerted by the concentration polarisation

effect. A comparison of the difference in normalised mass fraction for each ion

between the constant rejection case and variable rejection case is shown in Figure

280
7.21 and Figure 7.22 for Rech = 100 and Rech = 200, respectively. The variation is

relatively small (in the order of one percent of the normalised mass fraction).

Na+

Cl-

SO42-

Figure 7.21 – Difference in predicted ionic mass fraction distribution for spacer-filled channel between

constant rejection model and variable rejection model for multiple salt solution (NaCl:Na2SO4 with molar

ratio 0.2:0.8, overall ionic strength 34.1 eq m-3), Rech = 100

Some brief comments can be made about the differences in predicted ionic

concentration fields shown in Figure 7.21 and Figure 7.22. The main region where

differences are observed is in the vicinity of the filaments, which is unsurprising

considering the relatively rapid variation in rejection predicted by the coupled

model for these regions. Mass fractions predicted by the variable rejection model

281
are generally lower in the vicinity of these recirculation zones for the positively

rejected ions, and higher for the negatively rejected Cl- ion.

A small difference is also apparent along the remainder of the membrane surface

outside the filament recirculation zones: the mass fractions predicted by the

variable rejection model for the positively rejected ions are generally greater in

the area opposing the filaments, and lesser in the remaining area adjacent to the

filaments. The opposite of this trend is apparent for the negatively rejected Cl- ion.

This trend is more apparent for the higher channel Reynolds number, and for the

Cl- and SO42- ions. The overall difference is also greater for the higher channel

Reynolds number, which corresponds to the increasing hydrodynamic influence of

the filament recirculation zones.

282
Na+

Cl-

SO42-

Figure 7.22 – Difference in predicted ionic mass fraction distribution for spacer-filled channel between

constant rejection model and variable rejection model for multiple salt solution (NaCl:Na2SO4 with molar

ratio 0.2:0.8, overall ionic strength 34.1 eq m-3), Rech = 200

However, this variation in the ionic concentration field is not significant enough

to cause significant variation in the computed velocity and pressure fields between

the constant rejection and variable rejection solutions. It does, however, impact to

a small degree on the distribution of the permeate flux at the membrane surface, as

shown in Figure 7.23 and Figure 7.24 for Rech = 100 and Rech = 200, respectively.

However, the overall effect on the permeate flux over the monitoring area is so

small as to be negligible (as would be expected when using constant rejections

283
derived from the area-averaged predictions of the coupled model). The variation

in wall shear between the cases is also negligible, and not shown.

Velocity streamlines

Flux JV (× 10-5 m³ m² s-1)

7.15
Lower wall

7.10 Upper wall

Lower wall constant rejection


7.05 Upper wall constant rejection

7.00

6.95

6.90

6.85

6.80

6.75

6.70

Figure 7.23 – Difference in predicted permeate flux within spacer-filled channel w between constant rejection

model and variable rejection model (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 100)

284
Velocity streamlines

Flux JV (× 10-5 m³ m² s-1)

6.90
Lower wall
Upper wall
6.80 Lower wall constant rejection
Upper wall constant rejection

6.70

6.60

6.50

6.40

6.30

6.20

6.10

Figure 7.24 – Difference in predicted permeate flux within spacer-filled channel w between constant rejection

model and variable rejection model (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

Rech = 200)

285
7.5 Coupled simulation of complex mixed solution in spacer filled

channel

We now perform a final set of simulations using the multiple salt solution

described in Section 7.4, but with the addition of an uncharged solute (sucrose) to

demonstrate the applicability of the model to describe mixtures of charged and

uncharged solutes. The salt inlet mass fractions are maintained as per previous

simulations to maintain a consistent ionic strength, but the additional sucrose

component is added at an inlet concentration of 0.001 kg kg-1.

The results indicate that the addition of the uncharged component (sucrose) has

several effects on the behaviour of the membrane channel, primarily in terms of

mass transfer through the membrane. The high sucrose concentration in the

concentration polarisation layer (seen in the predicted normalised mass fraction

distributions in Figure 7.25 and Figure 7.26 for Rech = 100 and 200, respectively)

has a retarding effect on both the permeate flux and rejection of charged solutes.

This is due to the high osmotic pressure exerted by the sucrose molecules, as well

as the correspondingly higher solution viscosity due to the presence of sucrose

within the concentration polarisation layer. (This trend was also observed

previously in simulations of the simple stirred cell arrangement in Section 6.4).

The average reduction in flux (values averaged over the monitoring area are

shown in Table 7.4) due to the addition of sucrose is approximately 9 %.

286
Na+

Cl-

SO42-

Sucrose

Figure 7.25 – Predicted ionic mass fraction distribution for spacer-filled channel using coupled-mass transfer

model for complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-
3
, and additional sucrose at 0.001 kg kg-1), Rech = 100

The presence of sucrose also slightly decreases the predicted magnitude of the

rejections of the ionic solutes (averaged solute rejections for the current case are

shown in Table 7.5). This is true for both the positively rejected ions and the

negatively rejected Cl- co-ion. However, this effect is smaller (in the order of one

to two percent), as previously observed in Section 6.4.

287
Na+

Cl-

SO42-

Sucrose

Figure 7.26 – Predicted ionic mass fraction distribution for spacer-filled channel using coupled-mass transfer

model for complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-
3
, and additional sucrose at 0.001 kg kg-1), Rech = 200

The uncharged solute addition however does not significantly alter the bulk

hydrodynamics of the channel, and the general flow features are very similar to

that observed previously. Representative plots of the predicted fluid streamlines

(Figure 7.27), velocity vectors (Figure 7.28), and vorticity fields (Figure 7.29) are

shown in the following.

288
Rech = 100

Rech = 200

Figure 7.27 – Predicted fluid streamlines for spacer-filled channel using coupled-mass transfer model for

complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1)

Rech = 100

Rech = 200

Figure 7.28 – Predicted velocity vectors for spacer-filled channel using coupled-mass transfer model for

complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1)

289
Rech = 100

Rech = 200

Figure 7.29 – Predicted vorticity fields for spacer-filled channel using coupled-mass transfer model for single

complex mixture solution (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and

additional sucrose at 0.001 kg kg-1)

Table 7.4 – Comparison of average permeate flux values for single salt (NaCl, overall ionic strength 34.1 eq

m-3), multiple salt (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3), and complex

mixture (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at

0.001 kg kg-1) cases

Solution Averaged permeate flux JV × 10-5 m3 m-2 s-1

Rech = 100 Rech = 200

Single salt (NaCl) 6.51 6.16

Multiple salt (NaCl:Na2SO4) 6.92 6.58

Complex mixture (NaCl:Na2SO4:sucrose) 6.31 6.00

The spatial variation in solute rejection, permeate flux, and wall shear are shown

in Figure 7.30 and Figure 7.31 for Rech = 100 and 200, respectively. Similar

trends are again observed for permeate flux and the rejection of positively rejected

290
solutes, with maximum values observed opposite the spacer filaments. A local

maximum is also again apparent at the reattachment point of the downstream

filament recirculation zone at Rech = 100, but disappears with the higher crossflow

velocity at Rech = 200. Minimum flux and rejection occur at the separation points

of the recirculation zones upstream and downstream of the filaments.

Conversely, for the negatively rejected Cl- co-ion, minimum values (i.e. the

negative rejection of the greatest magnitude) are observed opposite to the spacer

filaments. Maximum values (the smallest magnitude of the negative rejection) are

observed at the recirculation zone separation points upstream and downstream of

the filaments. A local minimum is also apparent at the lower channel Reynolds

number only for the reattachment point of the downstream filament recirculation

zone. The wall shear results, being largely a function of the bulk channel

hydrodynamics, remain essentially unchanged from previous model results.

Again considering a location midway between the fifth and sixth filaments, the

transverse variation in the normalised solute concentrations is shown in Figure

7.32 through Figure 7.35. The variation for the ionic solutes is similar to that

observed previously in Section 7.4, however, this variation is dwarfed in

magnitude by that of the additional sucrose. Wall concentrations for the sucrose

component are in the order of two to three times the bulk concentration, due to the

high rejection of sucrose (R § The variation is again more pronounced for

the lower channel Reynolds number due to the lower wall shear.

291
Velocity streamlines

Na+ rejection
0.205
Lower wall
0.200
Upper wall
0.195
0.190
0.185
0.180
Cl- rejection
-0.170
Lower wall
-0.180 Upper wall

-0.190
-0.200
-0.210
SO42- rejection
0.250
Lower wall
0.240 Upper wall

0.230
0.220
0.210
Sucrose rejection
0.770
Lower wall

0.765 Upper wall

0.760

0.755
Flux JV (× 10-5 m³ m² s-1)
7.00
Lower wall

6.50 Upper wall

6.00

5.50
Wall shear (Pa)
0.40
Lower wall
0.30 Upper wall

0.20
0.10
0.00

Figure 7.30 – Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl:Na2SO4:sucrose solution, Rech = 100)

292
Velocity streamlines

Na+ rejection
0.210
Lower wall

0.200 Upper wall

0.190

0.180
Cl- rejection
Lower wall
-0.180
Upper wall

-0.200

-0.220
SO42- rejection
0.250
Lower wall
0.240
Upper wall
0.230
0.220
0.210
0.200
Sucrose rejection
0.770
Lower wall
0.765
Upper wall
0.760
0.755
0.750
0.745
Flux JV (× 10-5 m³ m² s-1)
6.70 Lower wall

Upper wall
6.20

5.70

5.20
Wall shear (Pa)
1.00
Lower wall

Upper wall

0.50

0.00
Figure 7.31 - Predicted variation in mass transfer parameters within spacer-filled channel with spatially

varying permeate flux and solute rejection (NaCl:Na2SO4:sucrose solution, Rech = 200)

293
2.250

2.050 Cl-
Na+
Normalised mass fraction (Ȧi /ȦiB)

1.850 SO42-
Sucrose

1.650

1.450

1.250

1.050

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.32 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from lower membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 100)

2.050

Cl-
1.850
Na+
Normalised mass fraction (Ȧi /ȦiB)

SO42-
1.650 Sucrose

1.450

1.250

1.050

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.33 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from lower membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 200)

294
2.850

2.650
Cl-
2.450 Na+
Normalised mass fraction (Ȧi /ȦiB)
SO42-
2.250
Sucrose
2.050

1.850

1.650

1.450

1.250

1.050

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.34 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from upper membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 100)

3.350

Cl-
2.850 Na+
Normalised mass fraction (Ȧi /ȦiB)

SO42-
Sucrose
2.350

1.850

1.350

0.850
0.001 0.01 0.1
Dimensionless normal distance from membrane (|y|/2h)

Figure 7.35 - Predicted transverse variation in normalised solute concentration as a function of non-

dimensionalised distance from upper membrane wall within spacer-filled channel with spatially varying

permeate flux and solute rejection (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 200)

295
Finally, simulations were again made using fixed constant rejections to elucidate

the effect of the spatially varying rejection capability of the coupled approach. A

summary of the adopted constant rejection values derived from the averaged

rejections predicted by the coupled model is shown in Table 7.5.

Table 7.5 – Fixed rejection values used for uncoupled model comparison case (NaCl:Na2SO4 with molar ratio

0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1)

Component Rejection

Rech = 100 Rech = 200

Na+ 0.197 0.196

Cl- -0.197 -0.198

SO42- 0.229 0.228

Sucrose 0.763 0.759

As per the comparison in Section 7.4, the results do differ slightly between the

constant rejection case and the variable rejection (coupled) case. The difference in

normalised solute mass fractions is shown in Figure 7.36 and Figure 7.37 for Rech

= 100 and Rech = 200, respectively. The general trend is similar to that previously

observed, with predicted mass fractions for positively rejected solutes in the

variable rejection case being higher in the vicinity of the membrane opposite to

spacer filaments, and lower elsewhere in the vicinity of the membrane. In

particular, the comparative decrease in mass fraction is again most pronounced in

the recirculation zones immediately before and after the filaments. This effect is

more significant for the higher channel Reynolds number, and especially so for

the strongly rejected sucrose component (R § 7KLVLVGXHWRWKHDGGLWLRQDO

296
osmotic and viscous effects of the high sucrose concentration. In addition, a small

area of increased concentration is now observed directly after the fifth filament,

which is thought to represent a locally high concentration zone corresponding to

the separation point of the downstream filament recirculation zone. The converse

of these trends is observed for the negatively rejected Cl- co-ion, as previously

reported.

Na+

Cl-

SO42-

Sucrose

Figure 7.36 – Difference in predicted solute mass fraction distribution for spacer-filled channel between

constant rejection model and variable rejection model for complex mixture solution (NaCl:Na2SO4 with

molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1), Rech = 100

297
Na+

Cl-

SO42-

Sucrose

Figure 7.37 – Difference in predicted solute mass fraction distribution for spacer-filled channel between

constant rejection model and variable rejection model for complex mixture solution (NaCl:Na2SO4 with

molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3, and additional sucrose at 0.001 kg kg-1), Rech = 200

Once more, though, the bulk hydrodynamics, nor the overall permeate flux are not

significantly affected by this mass-transfer variation. However, a small variation

between the constant and variable rejection cases is again apparent, most notably

in the vicinity of the spacer filaments, as shown in Figure 7.38 and Figure 7.39.

298
Velocity streamlines

Flux JV (× 10-5 m³ m² s-1)


6.90
Lower wall

Upper wall
6.70 Lower wall constant rejection

Upper wall constant rejection

6.50

6.30

6.10

5.90

5.70

Figure 7.38 – Difference in predicted permeate flux within spacer-filled channel w between constant rejection

model and variable rejection model (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 100)

Velocity streamlines

Flux JV (× 10-5 m³ m² s-1)


6.80
Lower wall
Upper wall
6.60 Lower wall constant rejection
Upper wall constant rejection
6.40

6.20

6.00

5.80

5.60

5.40

5.20

5.00

Figure 7.39 – Difference in predicted permeate flux within spacer-filled channel w between constant rejection

model and variable rejection model (NaCl:Na2SO4 with molar ratio 0.2:0.8, overall ionic strength 34.1 eq m-3,

and additional sucrose at 0.001 kg kg-1, Rech = 200)

299
7.6 Conclusions

Using the coupled hydrodynamic – mass transfer model, a series of numerical

experiments have been carried out for a 2D spacer-filled membrane channel. The

more complex geometry of such a channel leads to more complex hydrodynamic

and mass transfer effects, and allows for a more realistic application of the

coupled model than the simplified verification cases presented in Chapter 6.

Simulation results showed that the coupled model was successfully able to

reproduce the general hydrodynamic behaviour of the membrane channel

predicted by an equivalent (non-coupled) constant rejection model. This included

effects such as recirculation zones located upstream and downstream of the spacer

filaments, and formation of the concentration polarisation boundary layer

(including an inverse concentration polarisation layer in the presence of

negatively rejected solutes).

However, the coupled model approach affords some additional scope for

predicting the mass-transfer behaviour in the vicinity of the membrane walls.

Examination of the permeate flux and solute rejection, both allowed to vary as a

function of the local hydrodynamic conditions in the couple model, showed some

predictable patterns. In general, the magnitude of both the permeate flux and

solute rejection:

300
x decrease on the downstream face of an adjacent filament, as the associated

recirculation zone separates from the membrane wall;

x increase towards the upstream side of an adjacent filament, as the previous

recirculation zone reattaches to the membrane wall; and

x at low channel Reynolds numbers, reach a local maximum slightly

downstream of the reattachment point.

These trends are consistent with previous observations in the literature that these

reattachment and separation points are associated with regions of high and low

mass transfer, respectively. At higher channel Reynolds numbers, the magnitudes

of the flux and solute rejections reach a maximum directly opposite the spacer

filaments, which is consistent with the present simulation results showing this

location to coincide with the maximum wall shear. The two channel Reynolds

numbers (Rech = 100, 200) chosen for these simulations thus appear to straddle the

point where the bulk hydrodynamic flow features, rather than the near-wall

recirculation regions, begin to govern the permeate flux behaviour for this

particular system.

The results also revealed the transverse gradient in solute concentration (when

normalised appropriately) to vary between individual solutes in a multicomponent

solution. Considering an axial location midway between filaments, this effect was

shown to be greater for the membrane wall opposing the downstream filament,

commensurate with the predicted trends in flux and rejection variation.

301
While the bulk hydrodynamic behaviour and overall permeate flux remained

essentially constant between cases using constant rejection and varying rejection

(i.e. the coupled model), some variation was observed in the near-wall region both

in terms of the solute mass fraction distribution, and the distribution of the

permeate flux. This variation became more apparent at higher Reynolds number,

and where higher solute rejections were predicted.

An interesting effect was also observed for the overall magnitude of permeate flux

when compared between a single salt solution, a multiple salt solution, and a

complex mixed solution (multiple salts with an additional uncharged solute).

Although constant ionic strength and hydrodynamic forcing parameters (inlet

velocity and transmembrane pressure) were maintained across all simulations,

significant variations in the overall magnitude of the permeate flux were

observed: a change in the salt composition from pure NaCl to a 0.2:08

NaCl:Na2SO4 mixture increased the overall flux by approximately six to seven

percent, while further adding sucrose to the salt mixture reduced the overall flux

by approximately nine percent. These results give some justification to the use of

such a coupled model, in that accounting for the subtle changes in the mass

transfer behaviour of the near-wall region. This evidently may have significant

effects on the flux behaviour of the membrane which are not intuitively apparent

without use of a rigorous simulation tool. Future numerical studies of this effect

using a wider range of solution compositions (i.e. a greater number of different

302
solutes, and range of ionic strengths) would be useful, and could be relatively

easily verified by experimental studies of the overall permeate flux magnitude.

In summary, the coupled model approach can be seen to be able to well describe

the behaviour of a membrane channel over a range of conditions. The principal

limitation of the method however remains its computational expense: as noted in

Chapter 6, the coupled method introduces at least an order of magnitude increase

in computational expense compared to a constant rejection model. In real terms, a

steady-state solution for one of the coupled models simulated in this chapter

requires in the order of several days to a week of computational time on a typical

desktop computer, depending on the complexity of the feed solution. This is for

most purposes prohibitively large to be used as a practical tool for rapidly

predicting membrane system behaviour, let alone considering the use of transient

models to describe the filtration history of a system. Nevertheless, the coupled

approach may prove to be a useful tool in examining the fundamental behaviour

of membrane channels, provided sufficiently detailed experimental data can be

gathered to test the model predictions; and, as computational abilities inevitably

increase, may become more widely accessible.

303
Chapter 8

Conclusions

8.1 Overview

This research has presented an integrated approach for predictive modelling of

membrane filtration, by coupling mass transfer models to predict removal of

pollutants by the membrane with computational fluid dynamics simulations of the

membrane channel hydrodynamics. The objective of the study to develop a

general predictive model of the membrane filtration process has been realised, at

least for the high-pressure processes of NF and RO. This has been achieved by the

use of rigorous multicomponent hydrodynamic models, including true permeable

wall boundary conditions allowing spatially variable permeate flux and solute

rejection, in conjunction with numerical methods for predicting rejection for

mixtures of arbitrary numbers of charged and uncharged solutes, including the

effects of molecular shape. The coupling of these traditionally disparate modelling

approaches has yielded a model which should more comprehensively describe the

complex behaviours in a membrane channel, including the common phenomenon

of concentration polarisation.

In the following, conclusions drawn from this research are presented first

regarding the use of computational fluid dynamics (CFD) techniques to describe

hydrodynamics in membrane filtration, and suggestions are made for future work

304
in this area. Secondly, concluding arguments are made on the use of mass transfer

models to predict rejection in membrane filtration, along with recommendations

for future extensions of the present work. Finally, a set of conclusions and

recommendations for future work using the coupled hydrodynamic-mass transfer

approach are presented, including potential future applications of the coupled

modelling approach.

8.2 Hydrodynamic modelling – conclusions and recommendations

The commercial CFD code Ansys CFX was employed to predict the basic

hydrodynamic behaviour of a range of membrane channel configurations typically

encountered in practical applications. This approach was shown to describe the

hydrodynamic behaviour of these channels well, and was verified against

analytical, experimental and computational data from the literature. The primary

features of the basic hydrodynamic simulations conducted were as follows:

x Multicomponent fluid models were successfully employed with

concentration-dependent, spatially variable transport properties. This

means that the relevant transport properties of the solution (the viscosity,

density, diffusivity, and osmotic pressure) can vary spatially within the

membrane channel. Importantly, this allows for the concentration

polarisation boundary layer common to high-pressure membrane filtration

processes to be simulated; and

305
x True permeable wall boundary conditions at the membrane surface were

imposed, which allows for a more realistic description of the

hydrodynamics near the membrane surface than the more simplistic

impermeable wall or dissolving-wall approximations traditionally used.

The other benefit of this approach is that, in conjunction with the spatially

variable transport properties of the multicomponent fluid approach, it

allows for the permeate flux to vary spatially throughout the membrane

channel, and hence provide a more physical description of the channel

hydrodynamics.

The resulting implementation is general enough to describe most practical

situations for the high pressure membrane filtration processes of NF and RO.

However, the methods used to determine the spatially varying transport properties

are derived from empirical relationships presented in the literature for a limited

range of compounds. Further research to develop more general methods for

predicting the variation in these properties would therefore be a significant

improvement.

In addition, the multicomponent fluid approach employed does not attempt to

describe the behaviour of solutions containing particulate components, as

observed in lower-pressure membrane filtration (MF and UF). Future extension of

the modelling approach to include this behaviour may be a worthwhile endeavour.

306
Laminar flow models were successfully used to predict membrane channel

hydrodynamics in the flow regimes typically observed for NF and RO

membranes. For higher channel Reynolds numbers, transient laminar flow models

with very fine spatial and temporal discretisation were shown to be effective at

directly simulating unsteady flow structures without recourse to turbulence

models. Turbulence models are generally unsuitable for the relatively low

Reynolds numbers encountered in NF and RO, though they may be applicable for

the higher velocities of MF and UF. Investigation of the application of turbulence

models for these low-pressure processes may provide a useful avenue of future

investigation.

Finally, simulations were performed for several simplified 2D geometries

representative of a range of practical membrane geometries. It was demonstrated

that these 2D simulations can describe the behaviour of membrane channels in

reasonable detail, both with and without hydrodynamic spacers in the channel.

Some investigation was carried out into the use of 3D models of membrane

channels. It was concluded that, given the currently available computational

resources, 3D modelling is still largely infeasible, especially considering the very

fine spatial discretisation required to resolve the fine concentration polarisation

boundary layer, and, in transient simulations of unsteady flow, the extremely

small temporal discretisation also required.

In summary, it is recommended that future investigation into the hydrodynamic

modelling of membrane channels focus on the following areas:

307
x Development of general methods for predicting spatial variation in

transport properties for arbitrary multicomponent mixtures;

x Incorporation of particulate transport effects to potentially describe UF

and MF processes;

x Study of turbulence modelling applications for UF and MF processes; and

x Use of 3D modelling as computational resources advance.

8.3 Mass transfer modelling for rejection prediction – conclusions and

recommendations

A refined version of the classic Donnan Steric Partitioning Model (DSPM) has

been developed to present a consistent approach for predicting the rejection of

both charged and uncharged solutes. The method is based on the numerical

solution to the extended Nernst-Planck equations, assuming that solute transport

through membrane pores occurs through the processes of convection, diffusion,

and electric migration. The membrane is characterised in terms of three physically

meaningful parameters: the pore radius, an effective pore length or membrane

thickness, and a volumetric charge density. The method presented here differs

from previous implementations in that:

x The influence of molecular shape on the steric partitioning of the solute is

considered for both charged and uncharged molecules. Solute molecule

308
geometry is represented by a characteristic length and width, rather than

assuming a spherical solute. Previous models have only employed

molecular shape effects for large uncharged solutes: here, this is extended

for all uncharged and charged solutes, with the exception of monatomic

ions (where the spherical assumption remains well justified);

x Numerical methods have been used to solve the system of non-linear

equations describing the Donnan partitioning at the pore inlet and outlet,

which have historically only been solved analytically for small numbers of

univalent solutes. The numerical method employed here allows the

solution to be generalised to arbitrary numbers of charged solutes of

arbitrary valence; and

x The numerical solution to the Donnan partioning problem is used in

conjunction with a numerical solution of the boundary value problem

posed by the extended Nernst-Planck equations to iteratively determine the

concentration profile of solutes throughout the membrane pores and hence

predict solute rejection. The general nature of the numerical solution

means that rejection may be predicted for complex solutions of more than

three charged solutes, which is uncommon in the literature. In addition, the

rejection of an arbitrary number of uncharged solutes may also be

predicted simultaneously.

309
The refined model was shown to well predict the rejection of solutions containing

uncharged compounds, single salt solutions, and multiple salt solutions. The

model was also able to predict rejections well for a variety of feed solutions for a

given membrane using a consistent set of fixed membrane parameters (pore radius

and membrane thickness) and an isotherm to describe the variation in the

membrane volumetric charge density with the ionic strength of the feed solution.

This is an improvement over similar previous model implementations by other

authors where these parameters are adjusted between different feed solutions to

provide an adequate fit. Instead, the fact that these parameters can be maintained

at constant values for a given membrane indicates that the membrane

characterisation method is robust, and the chosen parameters are indeed

physically representative of the membrane.

In some cases with large uncharged solutes, agreement between experimental data

and the refined model predictions was less than ideal. In particular, complete

rejection was predicted for some compounds where experimental data indicated

slightly less than unity rejection. This discrepancy is believed to be due to the

existence of a pore size distribution in the physical membrane which is not

accounted for in the model. Future work on extending the model should include

such a pore size distribution within the mathematical model, as well as developing

a method by which this distribution can be characterised using experimental data.

Finally, the refined DSPM presented here does not include dielectric effects,

which have relatively recently been shown to be important in predicting rejection

310
in the case of multivalent counter-ions. However, dielectric effects can be

successfully incorporated into a DSPM model, and the numerical methods

required for solution would be substantively similar to those presented here.

Therefore, it is recommended that future work to include dielectric exclusion

effects into the general numerical solution with molecular shape effects would be

a definite improvement in predicting solute rejection for these processes.

8.4 Coupled hydrodynamic-mass transfer modelling – conclusions and

recommendations

Using the hydrodynamic models of the membrane channel and the mass-transfer

models to describe solute rejection, a coupled modelling approach was developed

and successfully employed. Numerical routines were developed using Fortran to

interface with the CFX solver, and predict rejections for multicomponent

solutions to be applied as boundary conditions for the coupled hydrodynamic

model. The coupled model was successfully run for a range of conditions, and

indicated several interesting phenomena not commonly discussed in the

hydrodynamic modelling literature, including:

x Spatial variation in solute rejection over the membrane surface and along

the membrane channel;

311
x Solute-dependent concentration polarisation boundary layers (i.e. varying

concentration gradients in the near-wall region for different solutes in

multicomponent solutions);

x The formation of an inverse concentration polarisation boundary layer in

the presence of negatively rejected solutes; and

x Significant variation in the overall permeate flux magnitude for single-salt

solutions, multiple salt solutions, and complex mixtures (multiple salts and

uncharged solutes), even though consistent ionic strength was maintained

throughout the simulations.

The predicted effect on permeate flux magnitude can be ascribed to the coupled

model’s ability to more fully describe the variation in solution viscosity and

osmotic pressure in the near-wall region and concentration polarisation boundary

layer (which influence permeate flux to a significant extent in these processes). A

useful area for future investigation would be to carry out a more detailed

numerical study of this effect using a wider range of solution compositions, which

could be relatively easily verified by experimental studies of the overall permeate

flux magnitude.

Some additional conclusions can be made regarding the coupled model approach,

in particular the treatment of the diffusion of solutes within the bulk flow of the

hydrodynamic model. In the uncoupled hydrodynamic only approach, individual

diffusion coefficients could be employed for each solute; however, in the coupled

312
approach this could not guarantee electroneutrality conditions at the membrane

surface, necessary for the solution of the mass-transfer model to predict solute

rejection. A simplified approach using diffusivities of salt ion-pairs, rather than

individual ions, was successfully employed to enforce electroneutrality

throughout the hydrodynamic model. This avoids the severe numerical difficulties

encountered by incorporating a full electro-diffusion model (such as the extended

Nernst-Planck equation) into the bulk flow of the hydrodynamic model as

reported by other authors. However, overcoming these numerical difficulties to

provide a rigorous coupled model with global incorporation of electro-diffusive

effects, not just within the membrane pores, remains a pressing goal for this

modelling approach. Further research efforts would be well expended in this area.

Other recommendations for future work on a coupled hydrodynamic-mass transfer

model have largely been referred to in the previous sections on hydrodynamics

and mass transfer separately. These include:

x Development of a general method for prediction of mixed solution

transport properties;

x Extension of the models to 3D geometries if computational resources

allow;

x Incorporation of other effects to make the model more generally applicable

to the lower-pressure processes of MF and UF, such as turbulence

modelling and particulate modelling; and

313
x Refinement of the mass-transfer model to include dielectric exclusion

effects.

Along with the recommendations presented for future work shown above, it is

also useful to consider the potential future applications of the model. An area in

which the coupled model approach would be particularly useful is in the

modelling of fouling reduction techniques, such as pulsatile flow and use of

‘turbulence’ promoters (either traditional spacer nets or Dean vortex promoters).

The ability of the model to simulate the formation of the concentration

polarisation layer in detail lends itself well to evaluating the applicability of such

techniques, which are largely aimed at minimising or eliminating the

concentration polarisation phenomenon. In addition, the use of transient models to

simulate the most likely unsteady flow in such situations would be appropriate.

Extension of the modelling approach to include the physics of other fouling

reduction techniques (such as multiphase flow to describe air sparging, or acoustic

propagation to describe ultrasonic fouling reduction) may also provide a future

application of the coupled model approach.

Finally, while a general entirely predictive model of membrane filtration

processes may be an overly optimistic goal for the near future, it is possible that

computational resources will advance to the point where full 3D simulations of

entire membrane module are possible. In this case, the coupled hydrodynamic-

mass transfer modelling approach, with the appropriate extensions suggested here,

314
may finally realise the goal of a completely predictive model of membrane

filtration.

315
Appendix A

Sample MATLAB numerical codes for predictive rejection

model

A.1 Molecular length and width calculations

The following code is used to determine the molecular length and width of a

solute based on the Cartesian coordinate output from the MOPAC2009 software

using a three-dimensional bounding box method. A graphical sample of the code

output for five test molecules (glucose, sucrose, cyanazine, glycerine, and Vitamin

B12) is also shown.

function [L,MWd] = molprjn2(xyz,atoms)


% Calculate molecular length L and width MWd as per Kiso et al

% Lookup table for VDW radius


VDWlookup = {'Ag' 1.72
'Ar' 1.88
'As' 1.85
'Au' 1.66
'Br' 1.85
'C' 1.7
'Cd' 1.58
'Cl' 1.75
'Co' 2.00
'Cu' 1.4
'F' 1.47
'Ga' 1.87
'H' 1.2
'He' 1.4
'Hg' 1.55
'I' 1.98
'In' 1.93
'K' 2.75
'Kr' 2.02
'Li' 1.82
'Mg' 1.73

316
'N' 1.55
'Na' 2.27
'Ne' 1.54
'Ni' 1.63
'O' 1.52
'P' 1.8
'Pb' 2.02
'Pd' 1.63
'Pt' 1.72
'S' 1.8
'Se' 1.9
'Si' 2.1
'Sn' 2.17
'Te' 2.06
'Tl' 1.96
'U' 1.86
'Xe' 2.16
'Zn' 1.39};

% Loop over structure 'atoms' to look up VDW radius for each


atom
for i = 1:length(atoms)

% Loop through structure 'VDWlookup'


for j = 1:length(VDWlookup)

if isequal(atoms(i),VDWlookup(j,1))

VDW(i) = VDWlookup(j,2);

break

end

end %j

end %i

% Convert cell array to numeric array


VDW=cell2mat(VDW');

% Create cloud of n points at radius VDW around each point in


xyz
% Create n random angles
n = 100;
[randangle] = randomAngle3d(n);
clouds = [];
for i = 1:length(xyz)

for j = 1:length(randangle)

317
clouds = [clouds; xyz(i,:) +
VDW(i)*sph2cart2(randangle(j,:))];

end %j

end %i

% Find length L - very inefficient but it works...


L=0;
for i = 1:(length(clouds)-1)

for j = (i+1):length(clouds)

% Start at second atom


% Distance between atoms including VDW radius
distance=sqrt(sum((clouds(j,:)-clouds(i,:)).^2));

% Find greatest distance


if distance>L
L = distance;
first = i;
second = j;
end %if

end %j

end %i

% Find width W
W = 0;
for k = 1:(length(clouds)) % Check each point in clouds

% Point to test whether in cylinder or not


testpt = clouds(k,:);

% Endpoints of cylinder
pt1 = clouds(first,:);
pt2 = clouds(second,:);

% Length of cylinder squared


lengthsq = L^2;

% Translate so pt1 is origin


d = pt2 - pt1;

% Vector from pt1 to test point


p = testpt - pt1;

% Dot the d and pd vectors to see if point lies behind the


% cylinder cap at pt1.x, pt1.y, pt1.z
dotpd = dot(p,d);

318
% distance squared to the cylinder axis including VDW
radius:
dsq = dot(p,p) - dotpd^2/lengthsq;

if sqrt(dsq) > W
W = sqrt(dsq);
third = k;
end %if

end %k

% Create plane containing points first, second, third


plane = createPlane(clouds(first,:), clouds(second,:),
clouds(third,:));

% Transform projected points so that they all lie on global


planes
% Create transformation matrix from plane to global plane
Tform = createBasisTransform3d('global',plane);
% Determine angles of rotation of plane
[PHI, THETA, PSI] =
rotation3dToEulerAngles(createBasisTransform3d(plane,'global')
);
% Transform projected points onto plane
cloudsTform = transformPoint3d(clouds, Tform);

% Find bounding box on transformed plane


BOXTform = point3dBounds(cloudsTform);
% Get dimensions of bounding box
L1 = BOXTform(2)-BOXTform(1);
L2 = BOXTform(4)-BOXTform(3);
L3 = BOXTform(6)-BOXTform(5);
% Get centre of bounding box
C1 = 0.5*(BOXTform(1)+BOXTform(2));
C2 = 0.5*(BOXTform(3)+BOXTform(4));
C3 = 0.5*(BOXTform(5)+BOXTform(6));
boxcentreTform = [C1 C2 C3];

% Transform centre back to global plane


boxcentre = transformPoint3d(boxcentreTform,
createBasisTransform3d(plane,'global'));

% Get first, second and third atoms


firstxyz = floor(first / n);
secondxyz = floor(second / n);
thirdxyz = floor(third / n);

% Draw clouds, atom centres, and spheres


%drawPoint3d(clouds,'.')
hold on
for i = 1:length(xyz)

319
drawSphere(xyz(i,:),
0.5*ones(size(VDW(i))),'FaceAlpha',0.5,'FaceColor','r');
drawSphere(xyz(i,:), VDW(i),'FaceAlpha',0.1);
end

% Plot enclosing box


CUBOID = [boxcentre L1 L2 L3 PHI THETA PSI];
drawCuboid(CUBOID,'FaceAlpha',0.05)
drawPoint3d(boxcentre)
axis equal

% Calculate molecular width


S = L2*L;
MWd = 0.5*sqrt(S);

end % molprjn2

320
321
322
A.2 Non-spherical steric partition coefficient calculation

The following code is used to determine steric partitioning coefficients for non-

spherical solutes by numerical integration.

%% Steric hindrance calculator for non-spherical solute


% Based on model of Kiso et al. (2010)
% phi = kiso(rp,L,MWd)
% Arguments are:
%
% rp - membrane pore radius [m]
% L - molecular length including VDW radius [m]
% MWd - molecular width including VDW radius [m]
%
% Returns:
%
% phi - column vector of steric partition coefficients
%
%%

function phi = kiso(rp,L,MWd)

% Sub-function for evaluating overall solute partition


coefficient
function F = int(alpha) % alpha is angle of molecular
orientation [rad]
% Effective molecular length Lp [m]
Lp = L(j).*cos(alpha)+2.*MWd(j).*sin(alpha);
% Solute-accessible radius in the pore a [m]
a = (rp(j).^2-(Lp./2).^2).^0.5-MWd(j);
if isreal(a)==0
a = 0;
end
% Overall solute partition coefficient
F = (a./rp(j)).^2.*sin(alpha);
end %int

for j=1:length(L)
% Evaluate integral of function F over 0:pi/2
phi(j) = quad(@int,0,pi/2);
end %j

phi=phi';

end %kiso

323
A.3 Donnan partitioning solver

The following code is used to solve the Donnan partitioning problem for charged

solutes, using a numerical solver to find the roots of a system of non-linear

equations.

%% Donnan equilibrium constant root finder


% APPEARS TO WORK GENERALLY!!!!
% Calculates roots of expression for Donnan equilibrium
constant
%
% Syntax: K = doneqn3(z,C,zX,CX,phi)
%
% Arguments are:
%
% z - column vector of valences of ions 1 to n
% C - column vector of solution concentrations of ions i to n
[mol m^-3]
% zX - valence of fixed membrane charge
% CX - concentration of fixed membrane charge [mol m^-3]
% phi - column vector of steric partition coefficients

function K = doneqn3(z,C,zX,CX,phi)

% Electroneutrality in solution means sum(z.*C) = 0


% Electroneutrality in membrane means sum(z.*y) + zX*CX = 0
% where y is the vector of concentrations of ions i to n
INSIDE THE
% MEMBRANE [mol m^-3]
% We use the combination of these two conditions as the first
equation of
% our function F
% i.e. sum(z.*y) + zX*CX + sum(z.*C) = 0

% As Donnan potential is identical for each ion i etc:


% 1/z1*log(C1/(y1*phi1))=1/zi*log(Ci/(yi*phii))
% therefore for the jth ion:
% 1/z1*log(C1/(y1*phi1)) - 1/zj*log(Cj/(yj*phij)) = 0 for each
j, j ~=1:
% This will give us another (n-1) equations for our function
F, giving
% us n equations with n unknowns (y1,y2,...,yn)

% Nested subfunction for fsolve to solve


function F = obj(y)
F = zeros(length(z),1);
% j represents all ions besides the first

324
j = 2:length(z);
% Electroneutrality in membrane and solution
F(1) = sum(z.*y) + zX*CX + sum(z.*C); %sum(Jv*C.*(1-
rejection).*z)
% % Equality of Donnan potential for each ion
% F(j) = 1./z(1).*log(C(1)./(y(1).*phi(1))) -
1./z(j).*log(C(j)./(y(j).*phi(j)));
% Equality of Donnan potential for each ion
F(j) = 1./z(1).*log(y(1)./(C(1).*phi(1))) -
1./z(j).*log(y(j)./(C(j).*phi(j)));
end

zcheck = z/zX;

for i=1:length(zcheck)
if zcheck(i) < 0 % counter-ion
guess(i) = C(i)*abs(zcheck(i)*CX);
else % co-ion
guess(i) = C(i)*abs(zcheck(i)/CX);
end %if
end %i

% Set options for root-finding function fsolve


options = optimset('Display','off','TolFun',1e-15,'TolX',1e-
12);

% Solve using root-finding function fsolve with the initial


guess that
% y/C = guess(i) for each ion i
[x,fval] = fsolve(@obj,guess',options);

% Calculate K
K = C.*phi./real(x);

% Check all Donnan voltages are equal


Donnancheck = log(1./K)./z;

origx = x;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%
% fsolve may return negative roots which are not physically
meaningful
% Check to see if this is the case and if negative roots were
found solve
% again using random number multiplied by absolute value of
original
% solution as starting guess
while min(x)<0 || abs(max(Donnancheck) - min(Donnancheck)) >
1E-3

newx = abs(origx).*rand(1);

325
[x,fval] = fsolve(@obj,newx,options);

% Recalculate K
K = C.*phi./real(x);

% Recheck all Donnan voltages are equal


Donnancheck = log(1./K)./z;

end
%
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%
%

end

326
A.4 Nernst-Planck equation solver

The following code is used to solve the extended Nernst-Planck equations for

charged solutes, using a numerical solver to find the roots of a system of non-

linear equations.

%% Nernst-Planck equation solver


% FULLY WORKING
% Solves extended Nernst-Planck equation for arbitrary number of
ions in
% solution
%
% Syntax:
%
% rejection = np3n(C,z,D,rs,Jv,rp,deltax,zX,CX)
% rejection = np3n(C,z,D,rs,Jv,rp,deltax,zX,CX,steric)
% rejection = np3n(C,z,D,rs,Jv,rp,deltax,zX,CX,steric,guess)
%
% Arguments are:
%
% C - column vector of feed concentrations of ions [mol m^-3]
% (note 1 M = 1 mol L^-1 = 1000 mol m^-3)
% z - column vector of valences of ions
% D - column vector of diffusivities of ions [m^2 s^-1]
% rs - column vector of Stokes radii of ions [m]
% NOTE: if non-spherical solutes, specify rs = 0
% Jv - volume flux [m s^-1]
% rp - membrane pore radius [m]
% deltax - membrane pore length [m]
% zX - effective membrane charge valence
% CX - effective membrane charge density [equiv m^-3]
%
% Optional arguments are:
%
% steric - logical value. 0 - no steric effects, 1 - steric
effects.
% L - column vector of molecular lengths [m] - used if non-
spherical model
% desired
% MWd - column vector of molecular widths [m] - used if non-
spherical model
% desired
% NOTE: if spherical solutes, specify L = MWd = 0
% guess - vector of guesses for rejection of ions
%
% Calls the following user functions:
%

327
% doneqn3.m
% kiso.m (if non-spherical model desired)
%

function rejection =
np3n(C,z,D,rs,Jv,rp,deltax,zX,CX,steric,L,MWd,guess)

% Set steric effects to be included if not specified


if nargin < 10, steric = 1; end

% Check if molecular length and width parameters have been


specified
if nargin == 12, nonsph = 1; else nonsph = 0; end

% Set guess at rejections to 0.5 for all ions if not specified


if nargin <13, guess = 0.5*ones(size(C)); end

% tic

% Define physical constants


% Faraday constant [C mol^-1]
F = 96485.3365;
% Temperature [K]
T = 293;
% Ideal gas constant [J mol^-1 K^-1]
R = 8.3144621;

% Check input vectors are same length


if length(C) == length(z) && length(C) == length(z) &&
length(C) ==...
length(D) && length(C) == length(rs)
else
error('Input vectors must be same length!')
end

% Define rejections as unknown parameters to pass to solver


parameters = guess;

% Calculate ratio of Stokes radius to pore radius, lambda


lambda = rs/rp;

if steric == 1;
% Steric partioning is included
% If molecular length and width are specified, use non-
spherical pore
% model
if nonsph == 1;
for i = 1:length(rs)
if rs(i) == 0 % no radius specified, use non-
spherical model
phi(i,1) = kiso(rp,L(i),MWd(i));

328
% Estimate rs based on MWd to calculate
hindrance factors
% for non-spherical solute
rs(i) = (1.42*(MWd(i)*1E9)-0.142)*1E-9;;
else % radius specified, use spherical
model
phi(i,1) = (1 - lambda(i))^2;
end %if
end %i
else
% Calculate spherical steric partition factors for all
molecules
phi = (1 - lambda).^2;
end % if
else
% Steric partioning not included
% Leave all steric partition factors equal to one
phi = ones(size(C));
end

% Calculate hindrance factors for diffusion using relationship


% Bowen et al. (1996)
% Kd = -1.705.*lambda+0.946;
% Deen (1987)
Kd = 1-2.30*lambda+1.154*(lambda.^2)+0.224*(lambda.^3);

% Calculate hindrance factors for convection using


relationship
% Bowen et al. (1996))
% Kc = -0.301.*lambda+1.022;
% Deen (1987)
G = 1+0.054*lambda-0.988*(lambda.^2)+0.441*(lambda.^3);
Kc = (2-phi).*G;

% Calculate diffusivity in pore [m^2 s^-1]


Dp = D.*Kd;

% Calculate Donnan potential at inlet [V] which is constant


for all ions
% Use external function doneqn3 to calculate vector of inlet
Donnan
% voltages
Donnaninletvector =
R.*T./(z.*F).*log(1./doneqn3(z,C,zX,CX,phi));

% At the outlet the Donnan voltage will be given by:


% R.*T./(z.*F).*log(1./doneqn3(z,C.*(1-parameters),zX,CX,phi))
% This is incorporated into the boundary conditions as
parameters is as
% yet unknown

% Calculate concentrations just inside pore inlet

329
inlet = C.*phi.*exp(z.*F/(R*T).*Donnaninletvector);

% BVPINIT is used to form an initial guess for a mesh of 100


equally spaced
% points. The guesses for the concentrations within the pore
% are evaluated in NPINIT. A guess for unknown parameters (the
rejection
% of each ion) is the last argument of BVPINIT.
solinit = bvpinit(linspace(0,deltax,100),@npinit,parameters);

% The BVP solver returns the structure 'sol'. The computed


rejections are
% returned in the field sol.parameters.
sol = bvp4c(@npode,@npbc,solinit);

rejection = sol.parameters;

Donnanoutletvector = R.*T./(z.*F).*log(1./doneqn3(z,C.*(1-
rejection),zX,CX,phi));

% Check zero current condition


I = sum(Jv*C.*(1-rejection).*z);

% ------------------------------------------------------------
-----------
function dydx = npode(x,y,parameters)
% Derivative function. Variables provided by the outer
function
%dydx = Jv./Dp.*(Kc.*y-C.*phi.*(1-parameters))-
z.*y.*F/(R*T)*sum(Jv./Dp.*z.*(Kc.*y-C.*phi.*(1-
parameters)))/(F/(R*T)*sum(z.*z.*y));
%dydx = Jv./Dp.*(Kc.*y-C.*phi.*(1-parameters))-
z.*y.*sum(Jv./Dp.*z.*(Kc.*y-C.*phi.*(1-
parameters)))/(sum(z.*z.*y));
dydx = Jv./Dp.*(Kc.*y-C.*(1-parameters))-
z.*y.*sum(Jv./Dp.*z.*(Kc.*y-C.*(1-
parameters)))/(sum(z.*z.*y));
end
% ------------------------------------------------------------
function res = npbc(ya,yb,parameters)
% Boundary conditions. parameters is a required
argument.
res = [ ya - inlet
%yb - C.*phi.*(1-parameters).*exp(-
z.*F/(R*T).*R.*T./(z.*F).*(log(1./doneqn3(z,C.*phi.*(1-
parameters),zX,CX))))];
yb - C.*phi.*(1-
parameters).*exp((log(1./doneqn3(z,C.*(1-
parameters),zX,CX,phi))))];

end

330
% ------------------------------------------------------------
-------------
% Auxiliary function -- initial guess for the solution - OK
function yinit = npinit(x)
yinit = inlet.*(1-(x./deltax).*parameters);
end
% ------------------------------------------------------------
-------------

% toc

end % np3n

331
A.5 Uncharged solute rejection calculator

The following code is used to find the analytical solution for the rejection of

uncharged solutes.

%% Rejection calculator for uncharged solutes


%
% Syntax: rejection = unchargedJv(L,MWd,rp,Jv,deltax)
%
% Arguments are:
%
% L - molecular length including VDW radius [m]
% MWd - molecular width including VDW radius [m]
% rp - membrane pore radius [m]
% Jv - volumetric flux [m s^-1]
%
% Calls the following user functions:
%
% kiso.m
%

function rejection = unchargedJv(L,MWd,rp,Jv,deltax)

% Define physical constants


% Faraday constant [C mol^-1]
F = 96485.3365;
% Temperature [K]
T = 293;
% Ideal gas constant [J mol^-1 K^-1]
R = 8.3144621;
% Thickness of a water molecule [nm]
d = 0.28;
% Boltzmann constant k [J?K^?1]
k = 1.3806504E-23;
% Viscosity of bulk solution eta0 [Pa s]
eta0 = 0.001;

% Calculate steric partioning coefficient phi and molecular


width MWd by
% calling external function kiso
phi = kiso(rp,L,MWd);

% Calculate approximate Stokes radius


rs = (1.42*(MWd*1E9)-0.142)*1E-9;

% Calculate ratio of Stokes radius to pore radius, lambda


lambda = rs./rp;

332
% Calculate hindrance factors for diffusion using relationship
from Bowen
% et al. (1997)
Kd = 1-2.30*lambda+1.154*(lambda.^2)+0.224*(lambda.^3);

% Calculate lag coefficient using relationship from Bowen et


al. (1997)
G = 1+0.054*lambda-0.988*(lambda.^2)+0.441*(lambda.^3);

% Calculate hindrance factors for convection using


relationship from Bowen
% et al. (1997)
Kc = (2-phi).*G;

% Calculate viscosity of water in pore eta [Pa s]


eta = (1+18*(d*1E-9./rp)-9*(d*1E-9./rp).^2)*eta0;

% Estimate bulk diffusivity [m^2 s^-1]


D = (rs.*eta0).^-1*(k*T/(6*pi));

% Calculate diffusivity in pore [m^2 s^-1]


Dp = D.*Kd;

% Calculate Peclet number


Pe = Jv.*Kc.*deltax./Dp;

% Calculate rejection
rejection = 1-phi.*Kc./(1-(1-phi.*Kc).*exp(-Pe));

end % unchargedJv

333
Appendix B

Sample FORTRAN numerical codes for coupled rejection

model – User Routine approach

B.1 Wrapper subroutine

The following code is used to pass the arguments from the CFX solver to the

other subroutines used to perform the numerical calculations.

#include "cfx5ext.h"
dllexport(np3nwrapper)
SUBROUTINE NP3NWRAPPER (
& NLOC, NRET, NARG, RET, ARGS, CRESLT, CZ,DZ,IZ,LZ,RZ )
CC
CD User routine: template for user CEL function
CC
CC --------------------
CC Input
CC --------------------
CC
CC NLOC - size of current locale
CC NRET - number of components in result
CC NARG - number of arguments in call
CC ARGS() - (NLOC,NARG) argument values
CC
CC --------------------
CC Modified
CC --------------------
CC
CC Stacks possibly.
CC
CC --------------------
CC Output
CC --------------------
CC
CC RET() - (NLOC,NRET) return values
CC CRESLT - 'GOOD' for success
CC
CC --------------------

334
CC Details
CC --------------------
CC
CC
CC
CC============================================================
==========
C
C ------------------------------
C Preprocessor includes
C ------------------------------
C
C
C ------------------------------
C Global Parameters
C ------------------------------
C
C
C ------------------------------
C Argument list
C ------------------------------
C

INTEGER NLOC,NARG,NRET
C
CHARACTER CRESLT*(*)
C
REAL ARGS(NLOC,NARG), RET(NLOC,NRET)
C
INTEGER IZ(*)
CHARACTER CZ(*)*(1)
DOUBLE PRECISION DZ(*)
LOGICAL LZ(*)
REAL RZ(*)

C
C ----------------------------------------------------------
C Input arguments are:
C
C 'Na plus.mf' is stored in ARGS(1:NLOC,1)
C 'Cl minus.mf' is stored in ARGS(1:NLOC,2)
C 'Na plus.Molar Mass' is stored in ARGS(1:NLOC,3)
C 'Cl minus.Molar Mass' is stored in ARGS(1:NLOC,4)
C 'rsNa plus' is stored in ARGS(1:NLOC,5)
C 'rsCl minus' is stored in ARGS(1:NLOC,6)
C 'DNa plus' is stored in ARGS(1:NLOC,7)
C 'DCl minus' is stored in ARGS(1:NLOC,8)
C 'zNa plus' is stored in ARGS(1:NLOC,9)
C 'zCl minus' is stored in ARGS(1:NLOC,10)
C 'LNa plus' is stored in ARGS(1:NLOC,11)
C 'LCl minus' is stored in ARGS(1:NLOC,12)
C 'MWdNa plus' is stored in ARGS(1:NLOC,13)

335
C 'MWdCl minus' is stored in ARGS(1:NLOC,14)
C 'zX' is stored in ARGS(1:NLOC,15)
C 'CX' is stored in ARGS(1:NLOC,16)
C 'rp' is stored in ARGS(1:NLOC,17)
C 'deltax' is stored in ARGS(1:NLOC,18)
C 'FluxVar' is stored in ARGS(1:NLOC,19)
C
C Output arguments are:
C
C 'Rejection Na plus' is stored in RET(1:NLOC,1)
C ----------------------------------------------------------
C
C ------------------------------
C External routines
C ------------------------------
C
C
C ------------------------------
C Local Parameters
C ------------------------------
C
C
C ------------------------------
C Local Variables
C ------------------------------
C
REAL REJ
REAL ARGSOUT(NARG)
REAL REXX((NARG-5)/7)
INTEGER I,J

INTERFACE
SUBROUTINE NP3NVECTOR(ARGSOUT,NARG,REXX)
REAL :: ARGSOUT(:),REXX(:)
INTEGER NARG
END SUBROUTINE NP3NVECTOR
END INTERFACE

C
C ------------------------------
C Stack pointers
C ------------------------------
C
C=============================================================
==========
C
C ---------------------------
C Executable Statements
C ---------------------------
C
C------------------- Loop over locations NLOC ----------------
----------

336
DO I=1,NLOC

DO J = 1,NARG

ARGSOUT(J) = ARGS(I,J)

END DO

CALL NP3NVECTOR(ARGSOUT,NARG,REXX)

C Set rejection to value contained in REXX

RET(I,1) = REXX(1)

END DO

C------------------- End loop over locations NLOC ------------


----------

C Print results for last NLOC for debugging purposes

CALL USER_PRINT_REAL('Na plus.mf',ARGSOUT(1))


CALL USER_PRINT_REAL('Cl minus.mf',ARGSOUT(2))
CALL USER_PRINT_REAL('Na plus.Molar Mass',ARGSOUT(3))
CALL USER_PRINT_REAL('Cl minus.Molar Mass',ARGSOUT(4))
CALL USER_PRINT_REAL('rsNa plus',ARGSOUT(5))
CALL USER_PRINT_REAL('rsCl minus',ARGSOUT(6))
CALL USER_PRINT_REAL('DNa plus',ARGSOUT(7))
CALL USER_PRINT_REAL('DCl minus',ARGSOUT(8))
CALL USER_PRINT_REAL('zNa plus',ARGSOUT(9))
CALL USER_PRINT_REAL('zCl minus',ARGSOUT(10))
CALL USER_PRINT_REAL('LNa plus',ARGSOUT(11))
CALL USER_PRINT_REAL('LCl minus',ARGSOUT(12))
CALL USER_PRINT_REAL('MWdNa plus',ARGSOUT(13))
CALL USER_PRINT_REAL('MWdCl minus',ARGSOUT(14))
CALL USER_PRINT_REAL('zX',ARGSOUT(15))
CALL USER_PRINT_REAL('CX',ARGSOUT(16))
CALL USER_PRINT_REAL('rp',ARGSOUT(17))
CALL USER_PRINT_REAL('deltax',ARGSOUT(18))
CALL USER_PRINT_REAL('FluxVar',ARGSOUT(19))

CALL USER_PRINT_REAL('Rejection',REXX(1))
CALL USER_PRINT_REAL('Rejection',REXX(2))

C
C Set success flag.
CRESLT = 'GOOD'
C
C=============================================================
==========

337
END

338
B.2 Steric partition subroutine

The following code is used to determine steric partitioning coefficients for

spherical or non-spherical solutes by numerical integration.

subroutine kiso(rp,L,MWd,rs,phi)
! uses QAG (included in quadpack.f90) to find the steric
partioning
! coefficient by numerical integration

implicit none

real, parameter :: llim = 0.0E+00


real abserr
real rlim
real, parameter :: epsabs = 0.0E+00
real, parameter :: epsrel = 0.001E+00
integer ier
integer, parameter :: key = 6
integer neval
real, parameter :: pi = 3.141592653589793E+00
real result
real :: rp,L(:),MWd(:),rs(:),phi(:)
integer j
real Lj,MWdj

rlim = pi/2.e0

do j = 1,size(L)

Lj = L(j)
MWdj = MWd(j)

! If no molecular dimensions supplied, assume


spherical
if (Lj .eq. 0) then ! spherical

phi(j) = (1-(rs(j)/rp))**2

else ! non-spherical

call qag ( phif, llim, rlim, epsabs, epsrel,


key, result, abserr, neval, ier )

phi(j) = result

rs(j) = (1.42*(MWd(j)*1E9)-0.142)*1E-9

339
end if

end do

contains

function phif ( alpha )


! phif is the integrand function

real :: phif,alpha,Lp,a

! Effective molecular length Lp [m]


Lp = Lj*cos(alpha)+2*MWdj*sin(alpha)

! Solute-accessible radius in the pore a [m]


if (rp<(Lp/2)) then
a = 0
else
a = (rp**2-(Lp/2)**2)**0.5-MWdj
end if

! Overall solute partition coefficient


phif = (a/rp)**2*sin(alpha)

return

end function phif

end subroutine kiso

340
B.3 Donnan partitioning solver

The following code is used to solve the Donnan partitioning problem for charged

solutes, using a numerical solver to find the roots of a system of non-linear

equations.

subroutine doneqn3(z,C,zX,CX,phi,K)
! uses HYBRD1 (included in minpack.f90) to solve Donnan
partitioning problem

implicit none

integer ( kind = 4 ) n

real ( kind = 8 ), allocatable :: fvec(:),y(:),origy(:)


integer ( kind = 4 ) :: iflag,info,i
real ( kind = 8 ) tol
real :: z(:),C(:),zX,CX,phi(:),K(:)
real, allocatable :: zcheck(:),Donnancheck(:)
real :: miny,diffDonnancheck,harvest,Donnantol

interface
subroutine hybrd1 ( f02, n, y, fvec, tol, info )
integer ( kind = 4 ) n
real ( kind = 8 ) y(n)
real ( kind = 8 ) fvec(n)
real ( kind = 8 ) tol
integer ( kind = 4 ) info
external f02
end subroutine hybrd1
end interface

n = size(C)

allocate(fvec(n))
allocate(y(n))
allocate(origy(n))
allocate(zcheck(n))
allocate(Donnancheck(n))

zcheck = z/zX

! Initial guess for solution


do i = 1, n

if (zcheck(i) < 0) then !counter-


ion

341
y(i) = C(i)*abs(zcheck(i)*CX)
else !co-ion
y(i) = C(i)*abs(zcheck(i)/CX)
end if

end do

K = y

! Set options for hybrd1


iflag = 1
tol = 1.0D-6

! Call hybrd1 to solve equations


call hybrd1 ( f02, n, y, fvec, tol, info)

! Calculate K
K = C*phi/abs(y);

! Tolerance for Donnan voltage check


Donnantol = 1E-3

! Check all Donnan voltages are equal


Donnancheck = log(1.0/K)/z;

miny = minval(abs(y))
diffDonnancheck = abs(maxval(Donnancheck) -
minval(Donnancheck))

if ( diffDonnancheck>Donnantol ) then

origy = y
call random_seed

end if

! hybrd1 may not return the correct solution


! Check to see if this is the case and if so solve
! again using random number multiplied by absolute value
of original
! solution as starting guess
do while ((miny .lt. 0.0) .or.
(diffDonnancheck>Donnantol))

call random_number(harvest)

y = abs(origy)*harvest;

! Call hybrd1 to solve equations


call hybrd1 ( f02, n, y, fvec, tol, info)

! Recalculate K

342
K = C*phi/abs(y);

! Recheck all Donnan voltages are equal


Donnancheck = log(1.0/K)/z;

miny = minval(y)
diffDonnancheck = abs(maxval(Donnancheck) -
minval(Donnancheck))

end do

deallocate(fvec)
deallocate(y)
deallocate(origy)
deallocate(zcheck)
deallocate(Donnancheck)

contains

subroutine f02 (n, y, fvec, iflag)

implicit none

integer ( kind = 4 ) n
real ( kind = 8 ) y(n)
real ( kind = 8 ) fvec(n)
integer ( kind = 4 ) iflag
integer ( kind = 4 ) j

! Electroneutrality in solution and membrane


fvec(1) = sum(z*abs(y)) + zX*CX + sum(z*C)

! Equality of Donnan potential:


do j = 2, n
fvec(j) = (1.0/z(1))*log(abs(y(1))/(C(1)*phi(1)))
- (1.0/z(j))*log(abs(y(j))/(C(j)*phi(j)))
end do

end subroutine f02

end subroutine doneqn3

343
B.4 Nernst-Planck equation solver

The following code is used to solve the extended Nernst-Planck equations for

charged solutes, using a numerical solver to find the roots of a system of non-

linear equations.

subroutine np3nvector(ARGSOUT,NARG,rejection)

use BVP_M

implicit none
TYPE(BVP_SOL) :: SOL

INTEGER NARG
REAL :: ARGSOUT(:),rejection(:)

INTEGER NODE,NPAR,LEFTBC,RIGHTBC

! Define input arguments from vector ARGSOUT


real,allocatable ::
L(:),MWd(:),rs(:),D(:),z(:),C(:),mf(:),mm(:)
real rp,deltax,zX,CX,Jv

! Define physical constants


real, parameter :: F = 96485.3365E0
real, parameter :: T = 293.0E0
real, parameter :: R = 8.3144621E0

real, allocatable :: phi(:),phiout(:),Kd(:),lambda(:)


real, allocatable ::
G(:),Kc(:),Dp(:),Donnaninletvector(:),Kin(:)
real, allocatable :: inlet(:),Kout(:),Donnanoutletvector(:)
real Icheck
double precision :: xmesh(10)
double precision, allocatable :: rejectionguess(:)
integer j

! Interface block - necessary to ensure arrays are passed


! in Fortran 90 style, rather than Fortran 77 -
! see http://w3.pppl.gov/~hammett/comp/f90_arrays.txt
interface
subroutine kiso(rp,L,MWd,rs,phi)
real :: rp,L(:),MWd(:),rs(:),phi(:)
end subroutine kiso
subroutine doneqn3(z,C,zX,CX,phi,K)
real :: z(:),C(:),zX,CX,phi(:),K(:)
end subroutine doneqn3

344
end interface

! Define global variables for the number of differential


! equations, NODE, the number of unknown parameters, NPAR,
! and the number of boundary conditions at the left end of
! the interval, LEFTBC. The number of boundary conditions
! at the right end is NODE+NPAR-LEFTBC.
! For this particular problem:
! NODE = NPAR = LEFTBC = RIGHTBC = (NARG-5)/7, which is the
! number of ions present
!INTEGER, PARAMETER :: NODE=2
!INTEGER, PARAMETER :: NPAR=NODE
!INTEGER, PARAMETER :: LEFTBC=NODE
!INTEGER, PARAMETER :: RIGHTBC=NODE
NODE = (NARG-5)/7
NPAR=NODE
LEFTBC=NODE
RIGHTBC=NODE

! Allocate vectors
allocate(L(NODE),MWd(NODE),rs(NODE),D(NODE),z(NODE),C(NODE),mf
(NODE),mm(NODE))
allocate(phi(NODE),phiout(NODE),Kd(NODE),lambda(NODE))
allocate(G(NODE),Kc(NODE),Dp(NODE),Donnaninletvector(NODE),Kin
(NODE))
allocate(inlet(NODE),Kout(NODE),Donnanoutletvector(NODE))
allocate(rejectionguess(NODE))

! Get input arguments from vector ARGSOUT


do j = 1,NODE

mf(j) = ARGSOUT(j)
mm(j) = ARGSOUT(1*NODE+j)
rs(j) = ARGSOUT(2*NODE+j)
D(j) = ARGSOUT(3*NODE+j)
z(j) = ARGSOUT(4*NODE+j)
L(j) = ARGSOUT(5*NODE+j)
MWd(j) = ARGSOUT(6*NODE+j)

end do

! First check if mass fractions are zero (first computional


step)
! If so, set rejection to zero and return
do j = 1,NODE

if (mf(j) .eq. 0.0) then

rejection = 0.0
return

end if

345
end do

Jv = ARGSOUT(size(ARGSOUT))
deltax = ARGSOUT(size(ARGSOUT)-1)
rp = ARGSOUT(size(ARGSOUT)-2)
CX = ARGSOUT(size(ARGSOUT)-3)
zX = ARGSOUT(size(ARGSOUT)-4)

! Calculate concentrations C from mass fractions mf and molar


masses mm
C = mf/mm*997.0*1000

! Use external subroutine kiso to calculate vector


! of steric partition coefficients phi and modify vector of
! Stokes radii rs as necessary
call kiso(rp,L,MWd,rs,phi)

! Calculate ratio of Stokes radius to pore radius, lambda


lambda = rs/rp

! Calculate hindrance factors for diffusion


Kd = 1-2.30*lambda+1.154*(lambda**2)+0.224*(lambda**3)

! Calculate hindrance factors for convection


G = 1+0.054*lambda-0.988*(lambda**2)+0.441*(lambda**3)
Kc = (2-phi)*G

! Calculate diffusivity in pore [m^2 s^-1]


Dp = D*Kd

! Calculate Donnan potential at inlet [V] which is constant


for all ions
! Use external subroutine doneqn3 to calculate vector of inlet
Donnan
! voltages
call doneqn3(z,C,zX,CX,phi,Kin)
Donnaninletvector = (R*T/(z*F))*log(1/Kin)

! Calculate concentrations just inside pore inlet


inlet = C*phi*exp(z*F/(R*T)*Donnaninletvector)

! Guess for rejections


rejectionguess = 0.5

rejection = rejectionguess

xmesh = BVP_LINSPACE(0.0D0,DBLE(deltax),10)

! BVP_INIT is used to form an initial guess for a mesh of 10


equally spaced

346
! points from zero to deltax (supplied in xmesh). The initial
guess is that
! of the inlet concentrations. A guess for unknown parameters
(rejectionguess)
! is supplied in an array as the last mandatory argument of
BVP_INIT.

SOL =
BVP_INIT(NODE,LEFTBC,xmesh,DBLE(inlet),rejectionguess,MAX_NUM_
SUBINTERVALS = 1000)

SOL = BVP_SOLVER(SOL,FSUB,BCSUB,TRACE=-1,TOL=1.0d-3,METHOD=4)

rejection = SOL%PARAMETERS

Donnanoutletvector = (R*T/(z*F))*log(1.0/Kout)

! Check zero current condition


Icheck = sum(Jv*C*(1-rejection)*z)

CALL BVP_TERMINATE(SOL)

! Dellocate vectors
deallocate(L,MWd,rs,D,z,C,mf,mm)
deallocate (phi,phiout,Kd,lambda)
deallocate (G,Kc,Dp,Donnaninletvector,Kin)
deallocate (inlet,Kout,Donnanoutletvector)
deallocate(rejectionguess)

contains

! System of ODEs (vector notation)


SUBROUTINE FSUB(X,Y,P,DYDX)
DOUBLE PRECISION :: X,Y(NODE),P(NPAR),DYDX(NODE)

dydx = Jv/Dp*(Kc*y-C*(1-P))-z*y*sum(Jv/Dp*z*(Kc*y-
C*(1-P)))/(sum(z*z*y));

END SUBROUTINE FSUB

! Boundary conditions for ODEs (vector notation)


SUBROUTINE BCSUB(YA,YB,P,BCA,BCB)
DOUBLE PRECISION :: YA(NODE),YB(NODE),P(NPAR),&

BCA(LEFTBC),BCB(RIGHTBC)

! First, return Donnan equilibrium coefficients at


outlet in Kout
call doneqn3(z,real(C*(1-P)),zX,CX,phi,Kout)

! Inlet boundary conditions

347
BCA = YA - inlet

! Outlet boundary conditions

BCB = YB - C*phi*(1.0D0-P)*exp(log(1.0D0/Kout))

END SUBROUTINE BCSUB

end subroutine np3nvector

348
Bibliography

[1] M.C. Porter, Handbook of industrial membrane technology, in, Noyes Publications,
New Jersey, 1990.

[2] M. Mulder, Basic principles of membrane technology, Kluwer Academic, Dordrecht,


Netherlands, 1991.

[3] A.I. Schäfer, Natural organics removal using membranes : principles, performance
and cost, Technomic Publishing, Lancaster, Pa., 2001.

[4] A.E. Childress, M. Elimelech, Relating Nanofiltration Membrane Performance to


Membrane Charge (Electrokinetic) Characteristics, Environ. Sci. Technol., 34 (2000)
3710-3716.

[5] E. Pellerin, E. Michelitsch, K. Darcovich, S. Lin, C.M. Tam, Turbulent transport in


membrane modules by CFD simulation in two dimensions, J. Membrane Sci., 100 (1995)
139-153.

[6] J.M. Benito, I.U. Richard, Modeling Concentration-Polarization in Reverse Osmosis


Spiral-Wound Elements, J. Environ. Eng., 122 (1996) 292-298.

[7] S. Kim, E.M.V. Hoek, Modeling concentration polarization in reverse osmosis


processes, Desalination, 186 (2005) 111-128.

[8] S.S. Sablani, M.F.A. Goosen, R. Al-Belushi, M. Wilf, Concentration polarization in


ultrafiltration and reverse osmosis: a critical review, Desalination, 141 (2001) 269-289.

[9] J.G. Wijmans, S. Nakao, J.W.A. Van Den Berg, F.R. Troelstra, C.A. Smolders,
Hydrodynamic resistance of concentration polarization boundary layers in ultrafiltration,
J. Membrane Sci., 22 (1985) 117-135.

[10] M. Al-Ahmad, F.A. Abdul Aleem, A. Mutiri, A. Ubaisy, Biofouling in RO


membrane systems Part 1: Fundamentals and control, Desalination, 132 (2000) 173-179.

[11] A. Al-Amoudi, R.W. Lovitt, Fouling strategies and the cleaning system of NF
membranes and factors affecting cleaning efficiency, J. Membrane Sci., 303 (2007) 4-28.

[12] S.n.F.E. Boerlage, Scaling and particulate fouling in membrane filtration systems, in,
Swets & Zeitlinger, Lisse, 2001.

349
[13] J.S. Baker, L.Y. Dudley, Biofouling in membrane systems -- A review, Desalination,
118 (1998) 81-89.

[14] T.V. Knyazkova, A.A. Maynarovich, Recognition of membrane fouling: testing of


theoretical approaches with data on NF of salt solutions containing a low molecular
weight surfactant as a foulant, Desalination, 126 (1999) 163-169.

[15] N. Hilal, O.O. Ogunbiyi, N.J. Miles, R. Nigmatullin, Methods Employed for Control
of Fouling in MF and UF Membranes: A Comprehensive Review, Separation Science and
Technology, 40 (2005) 1957 - 2005.

[16] W. Mohammad, L. Pei, A. Kadhum, Characterization and identification of rejection


mechanisms in nanofiltration membranes using extended Nernst–Planck model, Clean
Technologies and Environmental Policy, 4 (2002) 151-156.

[17] W.R. Bowen, H. Mukhtar, Characterisation and prediction of separation performance


of nanofiltration membranes, J. Membrane Sci., 112 (1996) 263-274.

[18] S. Bandini, D. Vezzani, Nanofiltration modeling: the role of dielectric exclusion in


membrane characterization, Chemical Engineering Science, 58 (2003) 3303-3326.

[19] A. Szymczyk, P. Fievet, Investigating transport properties of nanofiltration


membranes by means of a steric, electric and dielectric exclusion model, J. Membrane
Sci., 252 (2005) 77-88.

[20] L. Heng, Y. Yanling, G. Weijia, L. Xing, L. Guibai, Effect of pretreatment by


permanganate/chlorine on algae fouling control for ultrafiltration (UF) membrane system,
Desalination, 222 (2008) 74-80.

[21] A. Maartens, P. Swart, E.P. Jacobs, Feed-water pretreatment: methods to reduce


membrane fouling by natural organic matter, J. Membrane Sci., 163 (1999) 51-62.

[22] Y. Chen, B.Z. Dong, N.Y. Gao, J.C. Fan, Effect of coagulation pretreatment on
fouling of an ultrafiltration membrane, Desalination, 204 (2007) 181-188.

[23] W. Ma, Y. Zhao, L. Wang, The pretreatment with enhanced coagulation and a UF
membrane for seawater desalination with reverse osmosis, Desalination, 203 (2007) 256-
259.

[24] R. Ghosh, Z.F. Cui, Fractionation of BSA and lysozyme using ultrafiltration: effect
of pH and membrane pretreatment, J. Membrane Sci., 139 (1998) 17-28.

350
[25] L. Masse, D.I. Massé, Y. Pellerin, The effect of pH on the separation of manure
nutrients with reverse osmosis membranes, J. Membrane Sci., 325 (2008) 914-919.

[26] M.M.T. Khan, W. Jones, A. Camper, S. Takizawa, H. Katayama, F. Kurisu, S.


Ohgaki, Powdered activated carbon and biofiltration improve MF performance: Part I,
Membrane Technology, 2007 (2007) 7-10.

[27] H.K. Shon, S. Vigneswaran, J. Cho, Comparison of physico-chemical pretreatment


methods to seawater reverse osmosis: Detailed analyses of molecular weight distribution
of organic matter in initial stage, J. Membrane Sci., 320 (2008) 151-158.

[28] M. Gryta, Chemical pretreatment of feed water for membrane distillation, Chemical
Papers, 62 (2008) 100-105.

[29] S.L. Kim, J. Paul Chen, Y.P. Ting, Study on feed pretreatment for membrane
filtration of secondary effluent, Separation and Purification Technology, 29 (2002) 171-
179.

[30] B. Van der Bruggen, D. Segers, C. Vandecasteele, L. Braeken, A. Volodin, C. Van


Haesendonck, How a Microfiltration Pretreatment Affects the Performance in
Nanofiltration, Separation Science and Technology, 39 (2005) 1443 - 1459.

[31] C.K. Teng, M.N.A. Hawlader, A. Malek, An experiment with different pretreatment
methods, Desalination, 156 (2003) 51-58.

[32] J. Xu, G. Ruan, X. Chu, Y. Yao, B. Su, C. Gao, A pilot study of UF pretreatment
without any chemicals for SWRO desalination in China, Desalination, 207 (2007) 216-
226.

[33] G.K. Pearce, The case for UF/MF pretreatment to RO in seawater applications,
Desalination, 203 (2007) 286-295.

[34] H. Ma, C.N. Bowman, R.H. Davis, Membrane fouling reduction by backpulsing and
surface modification, J. Membrane Sci., 173 (2000) 191-200.

[35] R. Ghidossi, D. Veyret, P. Moulin, Computational fluid dynamics applied to


membranes: State of the art and opportunities, Chemical Engineering and Processing, 45
(2006) 437-454.

[36] J. Schwinge, D.E. Wiley, A.G. Fane, Novel spacer design improves observed flux, J.
Membrane Sci., 229 (2004) 53-61.

351
[37] V. Geraldes, V. Semião, M.N. de Pinho, The effect of the ladder-type spacers
configuration in NF spiral-wound modules on the concentration boundary layers
disruption, Desalination, 146 (2002) 187-194.

[38] V. Geraldes, V. Semião, M.N. de Pinho, Flow management in nanofiltration spiral


wound modules with ladder-type spacers, J. Membrane Sci., 203 (2002) 87-102.

[39] V. Geraldes, V. Semião, M. Norberta Pinho, Hydrodynamics and concentration


polarization in NF/RO spiral-wound modules with ladder-type spacers, Desalination, 157
(2003) 395-402.

[40] F. Li, G.W. Meindersma, A.B. de Haan, T. Reith, Optimization of non-woven


spacers by CFD and validation by experiments, Desalination, 146 (2002) 209-212.

[41] F. Li, W. Meindersma, A.B. de Haan, T. Reith, Experimental validation of CFD


mass transfer simulations in flat channels with non-woven net spacers, J. Membrane Sci.,
232 (2004) 19-30.

[42] C.P. Koutsou, S.G. Yiantsios, A.J. Karabelas, Numerical simulation of the flow in a
plane-channel containing a periodic array of cylindrical turbulence promoters, J.
Membrane Sci., 231 (2004) 81-90.

[43] V.V. Ranade, A. Kumar, Comparison of flow structures in spacer-filled flat and
annular channels, Desalination, 191 (2006) 236-244.

[44] J. Schwinge, D.E. Wiley, D.F. Fletcher, A CFD study of unsteady flow in narrow
spacer-filled channels for spiral-wound membrane modules, Desalination, 146 (2002)
195-201.

[45] M. Shakaib, S.M.F. Hasani, M. Mahmood, Study on the effects of spacer geometry
in membrane feed channels using three-dimensional computational flow modeling, J.
Membrane Sci., 297 (2007) 74-89.

[46] D.E. Wiley, D.F. Fletcher, Computational fluid dynamics modelling of flow and
permeation for pressure-driven membrane processes, Desalination, 145 (2002) 183-186.

[47] Y.-L. Li, K.-L. Tung, CFD simulation of fluid flow through spacer-filled membrane
module: selecting suitable cell types for periodic boundary conditions, Desalination, 233
(2008) 351-358.

[48] C. Cabassud, S. Laborie, J.M. Lainé, How slug flow can improve ultrafiltration flux
in organic hollow fibres, J. Membrane Sci., 128 (1997) 93-101.

352
[49] S.R. Bellara, Z.F. Cui, D.S. Pepper, Gas sparging to enhance permeate flux in
ultrafiltration using hollow fibre membranes, J. Membrane Sci., 121 (1996) 175-184.

[50] Z.F. Cui, K.I.T. Wright, Flux enhancements with gas sparging in downwards
crossflow ultrafiltration: performance and mechanism, J. Membrane Sci., 117 (1996) 109-
116.

[51] T. Taha, Z.F. Cui, CFD modelling of gas-sparged ultrafiltration in tubular


membranes, J. Membrane Sci., 210 (2002) 13-27.

[52] T. Taha, Z.F. Cui, Hydrodynamic analysis of upward slug flow in tubular
membranes, Desalination, 145 (2002) 179-182.

[53] Q.Y. Li, Z.F. Cui, D.S. Pepper, Effect of bubble size and frequency on the permeate
flux of gas sparged ultrafiltration with tubular membranes, Chemical Engineering
Journal, 67 (1997) 71-75.

[54] R. Ghosh, Z.F. Cui, Mass transfer in gas-sparged ultrafiltration: upward slug flow in
tubular membranes, J. Membrane Sci., 162 (1999) 91-102.

[55] S. Smith, T. Taha, Z. Cui, Enhancing hollow fibre ultrafiltration using slug-flow -- a
hydrodynamic study, Desalination, 146 (2002) 69-74.

[56] S.R. Smith, R.W. Field, Z.F. Cui, Predicting the performance of gas-sparged and
non-gas-sparged ultrafiltration, Desalination, 191 (2006) 376-385.

[57] B.J. Bellhouse, G. Costigan, K. Abhinava, A. Merry, The performance of helical


screw-thread inserts in tubular membranes, Separation and Purification Technology, 22-
23 (2001) 89-113.

[58] M.E. Brewster, K.-Y. Chung, G. Belfort, Dean vortices with wall flux in a curved
channel membrane system : 1. A new approach to membrane module design, J.
Membrane Sci., 81 (1993) 127-137.

[59] M. Bubolz, M. Wille, G. Langer, U. Werner, The use of dean vortices for crossflow
microfiltration: basic principles and further investigation, Separation and Purification
Technology, 26 (2002) 81-89.

[60] R. Moll, P. Moulin, D. Veyret, F. Charbit, Numerical simulation of Dean vortices:


fluid trajectories, J. Membrane Sci., 197 (2002) 157-172.

353
[61] P. Moulin, D. Veyret, F. Charbit, Dean vortices: comparison of numerical simulation
of shear stress and improvement of mass transfer in membrane processes at low
permeation fluxes, J. Membrane Sci., 183 (2001) 149-162.

[62] W. Wang, C.D. Bertram, D.E. Wiley, Effects of collapsible-tube-induced pulsation


vigour on membrane filtration performance, J. Membrane Sci., 288 (2007) 298-306.

[63] P. Czekaj, W. Mores, R.H. Davis, C. Güell, Infrasonic pulsing for foulant removal in
crossflow microfiltration, J. Membrane Sci., 180 (2000) 157-169.

[64] E. Spiazzi, J. Lenoir, A. Grangeon, A new generator of unsteady-state flow regime in


tubular membranes as an anti-fouling technique: A hydrodynamic approach, J. Membrane
Sci., 80 (1993) 49-57.

[65] P. Blanpain-Avet, N. Doubrovine, C. Lafforgue, M. Lalande, The effect of


oscillatory flow on crossflow microfiltration of beer in a tubular mineral membrane
system - Membrane fouling resistance decrease and energetic considerations, J.
Membrane Sci., 152 (1999) 151-174.

[66] H.-y. Li, Bertram, C.D., Wiley, D.E., Mechanisms by which pulsatile flow affects
cross-flow microfiltration, AIChE Journal, 44 (1998) 1950-1961.

[67] H.M. Kyllönen, P. Pirkonen, M. Nyström, Membrane filtration enhanced by


ultrasound: a review, Desalination, 181 (2005) 319-335.

[68] M.O. Lamminen, H.W. Walker, L.K. Weavers, Mechanisms and factors influencing
the ultrasonic cleaning of particle-fouled ceramic membranes, J. Membrane Sci., 237
(2004) 213-223.

[69] M.O. Lamminen, H.W. Walker, L.K. Weavers, Cleaning of particle-fouled


membranes during cross-flow filtration using an embedded ultrasonic transducer system,
J. Membrane Sci., 283 (2006) 225-232.

[70] X. Chai, T. Kobayashi, N. Fujii, Ultrasound-associated cleaning of polymeric


membranes for water treatment, Separation and Purification Technology, 15 (1999) 139-
146.

[71] S. Muthukumaran, S. Kentish, S. Lalchandani, M. Ashokkumar, R. Mawson, G.W.


Stevens, F. Grieser, The optimisation of ultrasonic cleaning procedures for dairy fouled
ultrafiltration membranes, Ultrasonics Sonochemistry, 12 (2005) 29-35.

[72] X. Chai, T. Kobayashi, N. Fujii, Ultrasound effect on cross-flow filtration of


polyacrylonitrile ultrafiltration membranes, J. Membrane Sci., 148 (1998) 129-135.

354
[73] T. Kobayashi, X. Chai, N. Fujii, Ultrasound enhanced cross-flow membrane
filtration, Separation and Purification Technology, 17 (1999) 31-40.

[74] I. Masselin, X. Chasseray, L. Durand-Bourlier, J.-M. Lainé, P.-Y. Syzaret, D.


Lemordant, Effect of sonication on polymeric membranes, J. Membrane Sci., 181 (2001)
213-220.

[75] R.-S. Juang, K.-H. Lin, Flux recovery in the ultrafiltration of suspended solutions
with ultrasound, J. Membrane Sci., 243 (2004) 115-124.

[76] E. Zondervan, B. Roffel, Evaluation of different cleaning agents used for cleaning
ultra filtration membranes fouled by surface water, J. Membrane Sci., 304 (2007) 40-49.

[77] B.R. Munson, D.F. Young, T.H. Okiishi, Fundamentals of fluid mechanics, 4th ed.
ed., Wiley, New York, 2002.

[78] M.B. Abbott, D.R. Basco, Computational fluid dynamics : an introduction for
engineers, Longman Scientific & Technical, Harlow, 1989.

[79] T.J. Chung, Computational fluid dynamics, Cambridge University Press, Cambridge,
U.K., 2002.

[80] J.H. Ferziger, M. Peric, Computational methods for fluid dynamics, Springer-Verlag,
New York, 1996.

[81] A.S. Berman, Laminar Flow in Channels with Porous Walls, J. Appl. Phys., 24
(1953) 1232-1235.

[82] M. Friedman, J. Gillis, Viscous flow in a pipe with absorbing walls, J. Appl. Mech.,
34 (1967) pp. 819–827.

[83] T. Mizushina, S. Takeshita, G. Unno, Study of flow in a porous tube with radial
mass flux, J. Chem. Eng. , 4 (1971) pp. 135–142.

[84] L.S. Galowin, M.J. De Santis, Investigation of laminar flow in a porous pipe with
variable wall suction, AIAA, 12 (1974) pp. 1585–1594.

[85] S.K. Karode, Laminar flow in channels with porous walls, revisited, J. Membrane
Sci., 191 (2001) 237-241.

[86] V. Nassehi, Modelling of combined Navier-Stokes and Darcy flows in crossflow


membrane filtration, Chemical Engineering Science, 53 (1998) 1253-1265.

355
[87] K. Damak, A. Ayadi, B. Zeghmati, P. Schmitz, A new Navier-Stokes and Darcy's
law combined model for fluid flow in crossflow filtration tubular membranes,
Desalination, 161 (2004) 67-77.

[88] P. Schmitz, M. Prat, 3-D Laminar stationary flow over a porous surface with suction:
Description at pore level, AIChE Journal, 41 (1995) 2212-2226.

[89] C. Gouverneur, Thèse de Doctorat, in, Institut National Polytechnique de Toulouse,


1991.

[90] L. Huang, M.T. Morrissey, Finite element analysis as a tool for crossflow membrane
filter simulation, J. Membrane Sci., 155 (1999) 19-30.

[91] C.J. Richardson, V. Nassehi, Finite element modelling of concentration profiles in


flow domains with curved porous boundaries, Chemical Engineering Science, 58 (2003)
2491-2503.

[92] S. Geissler, U. Werner, Dynamic model of crossflow microfiltration in flat-channel


systems under laminar flow conditions, Filtration & Separation, 32 (1995) 533-537.

[93] Y. Lee, M.M. Clark, Modeling of flux decline during crossflow ultrafiltration of
colloidal suspensions, J. Membrane Sci., 149 (1998) 181-202.

[94] V. Geraldes, V. Semião, M.N. de Pinho, Flow and mass transfer modelling of
nanofiltration, J. Membrane Sci., 191 (2001) 109-128.

[95] M. Niwa, H. Ohya, E. Kuwahara, Y. Negishi, Reverse osmotic concentration of


aqueous 2-butanone (methyl ethyl ketone), tetrahydrofuran and ethyl acetate solutions,
Journal of Chemical Engineering of Japan, 21 (1988) 164-171.

[96] S. Tanimura, S.-I. Nakao, S. Kimura, Transport analysis of reverse osmosis of


organic aqueous solutions, Journal of Chemical Engineering of Japan, 24 (1991) 364-371.

[97] C. Rosén, C. Trägårdh, Computer simulations of mass transfer in the concentration


boundary layer over ultrafiltration membranes, J. Membrane Sci., 85 (1993) 139-156.

[98] D.F. Fletcher, D.E. Wiley, A computational fluids dynamics study of buoyancy
effects in reverse osmosis, J. Membrane Sci., 245 (2004) 175-181.

[99] J.M. Miranda, J.B.L.M. Campos, Concentration polarization in a membrane placed


under an impinging jet confined by a conical wall -- a numerical approach, J. Membrane
Sci., 182 (2001) 257-270.

356
[100] G. Belfort, Fluid mechanics in membrane filtration: Recent developments, J.
Membrane Sci., 40 (1989) 123-147.

[101] V.V. Ranade, A. Kumar, Fluid dynamics of spacer filled rectangular and curvilinear
channels, J. Membrane Sci., 271 (2006) 1-15.

[102] G. Belfort, N. Nagata, Fluid mechanics and cross-flow filtration: some thoughts,
Desalination, 53 (1985) 57-79.

[103] G.A. Fimbres-Weihs, D.E. Wiley, Review of 3D CFD modeling of flow and mass
transfer in narrow spacer-filled channels in membrane modules, Chemical Engineering
and Processing: Process Intensification, 49 (2010) 759-781.

[104] B.K. Gupta, E.K. Levy, Symmetrical laminar channel flow with wall suction,
American Society of Mechanical Engineers, (1975).

[105] A.R. Da Costa, A.G. Fane, C.J.D. Fell, A.C.M. Franken, Optimal channel spacer
design for ultrafiltration, J. Membrane Sci., 62 (1991) 275-291.

[106] A.R. Da Costa, A.G. Fane, D.E. Wiley, Spacer characterization and pressure drop
modelling in spacer-filled channels for ultrafiltration, J. Membrane Sci., 87 (1994) 79-98.

[107] Z. Cao, D.E. Wiley, A.G. Fane, CFD simulations of net-type turbulence promoters
in a narrow channel, J. Membrane Sci., 185 (2001) 157-176.

[108] G. Sivashinsky, V. Yakhot, Negative viscosity effect in large-scale flows, Physics


of Fluids, 28 (1985) 1040-1042.

[109] Y. Miyake, K. Tsujimoto, H. Beppu, Direct numerical simulation of a turbulent


flow in a channel having periodic pressure gradient, International Journal of Heat and
Fluid Flow, 16 (1995) 333-340.

[110] J. Kim, P. Moin, R. Moser, Turbulence statistics in fully developed channel flow at
low Reynolds number, Journal of Fluid Mechanics Digital Archive, 177 (1987) 133-166.

[111] J. Schwinge, D.E. Wiley, D.F. Fletcher, Simulation of the Flow around Spacer
Filaments between Narrow Channel Walls. 1. Hydrodynamics, Industrial & Engineering
Chemistry Research, 41 (2002) 2977-2987.

[112] J. Schwinge, D.E. Wiley, D.F. Fletcher, Simulation of the Flow around Spacer
Filaments between Channel Walls. 2. Mass-Transfer Enhancement, Industrial &
Engineering Chemistry Research, 41 (2002) 4879-4888.

357
[113] G.A. Fimbres-Weihs, D.E. Wiley, Numerical study of two-dimensional multi-layer
spacer designs for minimum drag and maximum mass transfer, J. Membrane Sci., 325
(2008) 809-822.

[114] G.A. Fimbres-Weihs, D.E. Wiley, D.F. Fletcher, Unsteady Flows with Mass
Transfer in Narrow Zigzag Spacer-)LOOHG &KDQQHOVௗ $ 1XPHULFDO 6WXG\ ,QGXVWULDO 
Engineering Chemistry Research, 45 (2006) 6594-6603.

[115] S. Wardeh, H.P. Morvan, CFD simulations of flow and concentration polarization
in spacer-filled channels for application to water desalination, Chemical Engineering
Research and Design, 86 (2008) 1107-1116.

[116] S.K. Karode, A. Kumar, Flow visualization through spacer filled channels by
computational fluid dynamics I.: Pressure drop and shear rate calculations for flat sheet
geometry, J. Membrane Sci., 193 (2001) 69-84.

[117] C.P. Koutsou, S.G. Yiantsios, A.J. Karabelas, Direct numerical simulation of flow
in spacer-filled channels: Effect of spacer geometrical characteristics, J. Membrane Sci.,
291 (2007) 53-69.

[118] G.A. Fimbres-Weihs, D.E. Wiley, Numerical study of mass transfer in three-
dimensional spacer-filled narrow channels with steady flow, J. Membrane Sci., 306
(2007) 228-243.

[119] C.P. Koutsou, S.G. Yiantsios, A.J. Karabelas, A numerical and experimental study
of mass transfer in spacer-filled channels: Effects of spacer geometrical characteristics
and Schmidt number, J. Membrane Sci., 326 (2009) 234-251.

[120] F. Li, W. Meindersma, A.B. de Haan, T. Reith, Novel spacers for mass transfer
enhancement in membrane separations, J. Membrane Sci., 253 (2005) 1-12.

[121] J.L.C. Santos, V. Geraldes, S. Velizarov, J.G. Crespo, Investigation of flow patterns
and mass transfer in membrane module channels filled with flow-aligned spacers using
computational fluid dynamics (CFD), J. Membrane Sci., 305 (2007) 103-117.

[122] K.K. Lau, M.Z. Abu Bakar, A.L. Ahmad, T. Murugesan, Feed spacer mesh angle:
3D modeling, simulation and optimization based on unsteady hydrodynamic in spiral
wound membrane channel, J. Membrane Sci., 343 (2009) 16-33.

[123] ANSYS CFX-Solver Theory Guide, ANSYS, Inc., Canonsburg, PA, 2010.

[124] D. Bhattacharyya, S.L. Back, R.I. Kermode, M.C. Roco, Prediction of


concentration polarization and flux behavior in reverse osmosis by numerical analysis, J.
Membrane Sci., 48 (1990) 231-262.

358
[125] R. Krishna, J.A. Wesselingh, The Maxwell-Stefan approach to mass transfer,
Chemical Engineering Science, 52 (1997) 861-911.

[126] L. Onsager, THEORIES AND PROBLEMS OF LIQUID DIFFUSION, Annals of


the New York Academy of Sciences, 46 (1945) 241-265.

[127] S. Déon, P. Dutournié, L. Limousy, P. Bourseau, The two-dimensional pore and


polarization transport model to describe mixtures separation by nanofiltration: Model
validation, AIChE Journal, 57 (2011) 985-995.

[128] S. Bhattacharjee, J.C. Chen, M. Elimelech, Coupled model of concentration


polarization and pore transport in crossflow nanofiltration, AIChE Journal, 47 (2001)
2733-2745.

[129] J. Kucera, Reverse osmosis : design, processes, and applications for engineers,
Scrivener Pub. ;

Wiley, Salem, Mass.

Hoboken, N.J., 2010.

[130] P. Schausberger, N. Norazman, H. Li, V. Chen, A. Friedl, Simulation of protein


ultrafiltration using CFD: Comparison of concentration polarisation and fouling effects
with filtration and protein adsorption experiments, J. Membrane Sci., 337 (2009) 1-8.

[131] J.S. Vrouwenvelder, C. Picioreanu, J.C. Kruithof, M.C.M. van Loosdrecht,


Biofouling in spiral wound membrane systems: Three-dimensional CFD model based
evaluation of experimental data, J. Membrane Sci., 346 (2010) 71-85.

[132] M. Shakaib, S.M.F. Hasani, M. Mahmood, CFD modeling for flow and mass
transfer in spacer-obstructed membrane feed channels, J. Membrane Sci., 326 (2009) 270-
284.

[133] C. Picioreanu, J.S. Vrouwenvelder, M.C.M. van Loosdrecht, Three-dimensional


modeling of biofouling and fluid dynamics in feed spacer channels of membrane devices,
J. Membrane Sci., 345 (2009) 340-354.

[134] Y.-L. Li, K.-L. Tung, M.-Y. Lu, S.-H. Huang, Mitigating the curvature effect of the
spacer-filled channel in a spiral-wound membrane module, J. Membrane Sci., 329 (2009)
106-118.

[135] Y.-L. Li, K.-L. Tung, The effect of curvature of a spacer-filled channel on fluid
flow in spiral-wound membrane modules, J. Membrane Sci., 319 (2008) 286-297.

359
[136] S. Ma, L. Song, S.L. Ong, W.J. Ng, A 2-D streamline upwind Petrov/Galerkin
finite element model for concentration polarization in spiral wound reverse osmosis
modules, J. Membrane Sci., 244 (2004) 129-139.

[137] J. Schwinge, D.E. Wiley, D.F. Fletcher, Simulation of Unsteady Flow and Vortex
Shedding for Narrow Spacer-Filled Channels, Industrial & Engineering Chemistry
Research, 42 (2003) 4962-4977.

[138] K.S. Spiegler, O. Kedem, Thermodynamics of hyperfiltration (reverse osmosis):


criteria for efficient membranes, Desalination, 1 (1966) 311-326.

[139] O. Kedem, A. Katchalsky, Permeability of composite membranes. Part 1.-Electric


current, volume flow and flow of solute through membranes, Transactions of the Faraday
Society, 59 (1963) 1918-1930.

[140] B. Van der Bruggen, C. Vandecasteele, Modelling of the retention of uncharged


molecules with nanofiltration, Water Research, 36 (2002) 1360-1368.

[141] W.R. Bowen, A.W. Mohammad, Diafiltration by nanofiltration: Prediction and


optimization, AIChE Journal, 44 (1998) 1799-1812.

[142] J.M.M. Peeters, Characterization of nanofiltration membranes, in, Enschede, 1997,


pp. 159.

[143] J. Palmeri, P. Blanc, A. Larbot, P. David, Theory of pressure-driven transport of


neutral solutes and ions in porous ceramic nanofiltration membranes, J. Membrane Sci.,
160 (1999) 141-170.

[144] J. Schaep, C. Vandecasteele, A. Mohammad, W. Bowen, Analysis of the Salt


Retention of Nanofiltration Membranes Using the Donnan–Steric Partitioning Pore
Model, Separation Science & Technology, 34 (1999) 3009.

[145] J.M.M. Peeters, J.P. Boom, M.H.V. Mulder, H. Strathmann, Retention


measurements of nanofiltration membranes with electrolyte solutions, J. Membrane Sci.,
145 (1998) 199-209.

[146] C. Diaper, V. Correia, S. Judd, Characterisation of zirconium/poly(acrylic acid) low


pressure dynamically formed membranes by use of the extended Nernst-Planck equation,
J. Membrane Sci., 138 (1998) 135-140.

[147] X.-L. Wang, T. Tsuru, S.-i. Nakao, S. Kimura, The electrostatic and steric-
hindrance model for the transport of charged solutes through nanofiltration membranes, J.
Membrane Sci., 135 (1997) 19-32.

360
[148] W.R. Bowen, A.W. Mohammad, N. Hilal, Characterisation of nanofiltration
membranes for predictive purposes -- use of salts, uncharged solutes and atomic force
microscopy, J. Membrane Sci., 126 (1997) 91-105.

[149] W.R. Bowen, A.W. Mohammad, A theoretical basis for specifying nanofiltration
membranes -- Dye/salt/water streams, Desalination, 117 (1998) 257-264.

[150] W.R. Bowen, J.S. Welfoot, Modelling of membrane nanofiltration--pore size


distribution effects, Chemical Engineering Science, 57 (2002) 1393-1407.

[151] D. Vezzani, S. Bandini, Donnan equilibrium and dielectric exclusion for


characterization of nanofiltration membranes, Desalination, 149 (2002) 477-483.

[152] C. Labbez, P. Fievet, A. Szymczyk, A. Vidonne, A. Foissy, J. Pagetti, Retention of


mineral salts by a polyamide nanofiltration membrane, Separation and Purification
Technology, 30 (2003) 47-55.

[153] J. Schaep, C. Vandecasteele, A. Wahab Mohammad, W. Richard Bowen,


Modelling the retention of ionic components for different nanofiltration membranes,
Separation and Purification Technology, 22–23 (2001) 169-179.

[154] M.S. Hall, V.M. Starov, D.R. Lloyd, Reverse osmosis of multicomponent
electrolyte solutions Part I. Theoretical development, J. Membrane Sci., 128 (1997) 23-
37.

[155] W.M. Deen, Hindered transport of large molecules in liquid-filled pores, AIChE
Journal, 33 (1987) 1409-1425.

[156] B. Van der Bruggen, J. Schaep, W. Maes, D. Wilms, C. Vandecasteele,


Nanofiltration as a treatment method for the removal of pesticides from ground waters,
Desalination, 117 (1998) 139-147.

[157] B. Van der Bruggen, J. Schaep, D. Wilms, C. Vandecasteele, Influence of


molecular size, polarity and charge on the retention of organic molecules by
nanofiltration, J. Membrane Sci., 156 (1999) 29-41.

[158] Y. Kiso, T. Kon, T. Kitao, K. Nishimura, Rejection properties of alkyl phthalates


with nanofiltration membranes, J. Membrane Sci., 182 (2001) 205-214.

[159] Y. Kiso, K. Muroshige, T. Oguchi, M. Hirose, T. Ohara, T. Shintani, Pore radius


estimation based on organic solute molecular shape and effects of pressure on pore radius
for a reverse osmosis membrane, J. Membrane Sci., 369 (2011) 290-298.

361
[160] Y. Kiso, K. Muroshige, T. Oguchi, T. Yamada, M. Hhirose, T. Ohara, T. Shintani,
Effect of molecular shape on rejection of uncharged organic compounds by nanofiltration
membranes and on calculated pore radii, J. Membrane Sci., 358 (2010) 101-113.

[161] J.L.C. Santos, P. de Beukelaar, I.F.J. Vankelecom, S. Velizarov, J.G. Crespo, Effect
of solute geometry and orientation on the rejection of uncharged compounds by
nanofiltration, Separation and Purification Technology, 50 (2006) 122-131.

[162] M.J.D. Powell, A Fortran Subroutine for Solving Systems of Nonlinear Algebraic
Equations, in: P. Rabinowitz (Ed.) Numerical methods for nonlinear algebraic equations,
Gordon and Breach Science Publishers, 1970.

[163] L.F. Shampine, P.H. Muir, H. Xu, A User-Friendly Fortran BVP Solver, Promacon,
Athens, GRECE, 2006.

[164] L.F. Shampine, M.W. Reichelt, J. Kierzenka, Solving Boundary Value Problems
for Ordinary Differential Equations in MATLAB with bvp4c, in, 2000.

[165] MATLAB Release Notes, The MathWorks, Inc., 2009.

[166] J. Burkardt, Source Codes in Fortran 90, in.

[167] J. Stewart, Optimization of parameters for semiempirical methods V: Modification


of NDDO approximations and application to 70 elements, Journal of Molecular
Modeling, 13 (2007) 1173-1213.

[168] M.N. de Pinho, V. Semião, V. Geraldes, Integrated modeling of transport processes


in fluid/nanofiltration membrane systems, J. Membrane Sci., 206 (2002) 189-200.

362

También podría gustarte