978-3-030-81931-6 (2)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 749

Heinz Ludwig

Reverse Osmosis
Seawater
Desalination
Volume 1
Planning, Process Design and
Engineering – A Manual for Study and
Practice
Reverse Osmosis Seawater Desalination
Volume 1
Heinz Ludwig

Reverse Osmosis Seawater


Desalination Volume 1
Planning, Process Design and
Engineering – A Manual for Study and
Practice
Heinz Ludwig
Nufringen, Baden-Württemberg, Germany

ISBN 978-3-030-81930-9 ISBN 978-3-030-81931-6 (eBook)


https://doi.org/10.1007/978-3-030-81931-6

# The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface and Acknowledgement

During the more than five decades in which I have now been active in water,
wastewater, and environmental process engineering, from the 1970s onwards, mem-
brane technology, evolving from a novel and innovative process for separating solids
and dissolved salts from water, whose potential and applications were initially
perceived only as a possible complement to conventional solids separation, such
as filtration, sedimentation, and flotation, as well as desalination by ion exchange
and evaporative technologies has since advanced to become a dominant technology
for separating suspended and colloidal substances from water as well as for
desalination.
Following the initial phase of investigating the technical possibilities and
evaluating its economic potential in comparison with conventional treatment
technologies, the next phase in its development was quickly reached with diverse
applications in the industrial sector for treating process water, wastewater, and
process solutions. After the establishment of membrane technology within the
range of separation and desalination processes available and suitable for the treat-
ment of water and solutions, the next development step was its use in seawater and
brackish water desalination, initially in plants with a lesser product water capacity,
but then for large-scale plants alongside the thermal multi-stage flash (MSF) and
multi-effect distillation (MED) technologies that were already well established for
this application. In the past two decades, increasingly, it has found application in
stand-alone installations as an alternative to or in competition with evaporative
desalination technology, even to the extent of supplanting this altogether.
A particular attraction of working in a specific sector over many years is having
the opportunity to witness and participate in technological developments right from
their inception and up to the attainment of fully fledged maturity. In the case of
membrane technology, it is particularly its use in seawater desalination that has
developed during my professional career from the more industrially oriented small-
and medium-sized plants with a few 100 m³/day of treated water output to today’s
large-scale facilities for supplying drinking water with capacities of several
100,000 m³/day. It presented a professional challenge and was a particular motiva-
tion for me to play a part in introducing and establishing this technology for securing
the water supply of many cities and regions and even for parts of an entire continent
like Australia.

v
vi Preface and Acknowledgement

What has always fascinated me during my professional career in projects for


planning and realizing large seawater desalination plants is the diversity of
professions that come together in this sector in order to achieve the goal, namely
the design, erection, and operation of such a plant. This starts with the variety of
technical disciplines involved in the design and construction of its process engineer-
ing components. However, right from the project development stage, when decisions
have to be taken on strategies for the procurement, project implementation, and
contractual regulation of plant operation, economists, lawyers, and experts in con-
tract law play important roles. In addition, for major state-owned desalination
projects, media experts are oftentimes called in at this stage to help gain public
acceptance for the plant and its location. The success of a seawater desalination
project and the timing of its implementation is also greatly influenced by planning
and executing the procedures for obtaining permits under environmental regulations
and thus the involvement of ecologists with coordination of the various disciplines
that work with them as part of the plant’s planning and design teams. This multidis-
ciplinary approach to seawater desalination projects is therefore also reflected in the
structure and content of this book as described in Chap. 1.
The catalyst for writing this book was the presentation of the Jim Stewart Lifetime
Achievement Award to me by Global Water Intelligence (GWI) at the Global Water
Summit 2011 in Berlin, which honored my over 40 years of professional involve-
ment in the field of seawater desalination. As a follow-up to this acknowledgement
of my professional work, together with Fichtner GmbH & Co, KG, the engineering
consultancy company for which I was active for a large number of seawater desali-
nation projects over this entire period in a managerial capacity, as an expert and as a
technical advisor, the idea emerged of authoring a manual based on the experience I
had gained in this field, in which the state of the art of seawater desalination on the
basis of reverse osmosis technology is comprehensively presented, starting from its
physical, chemical, and process engineering fundamentals up to the planning and
design of these RO desalination plants.
The reference book that is the outcome of this collaboration is not an academic
textbook, but rather it is intended to serve as a practice-oriented introduction to the
planning processes and the approach for the design and realization of plants for the
application of membrane processes to seawater desalination and also to introduce the
relevant design and operating parameters. However, theoretical principles and the
findings of academic research that serve to provide an understanding of the
calculations involved in dimensioning the systems and components of such
installations are presented in detail and also clarified by means of the associated
mathematical algorithms.
This book sets out not only my own experiences and the knowledge that I have
acquired from my engineering and design activities but, especially in its sections
describing the design of the treatment processes of an SWRO plant as regards their
dimensioning and their operating characteristics, it is additionally based on volumi-
nous technical information from desalination plant constructors as well as from the
suppliers of membranes, conditioning chemicals, and the various process
components, together with results from calculation software provided by these
Preface and Acknowledgement vii

manufacturers for the application of their products. In addition, there are insights
gained from the exchange of information between the various disciplines within the
planning teams of an SWRO plant as well as during the phases for developing and
realizing such projects in dialogue with the various project participants.
Further contributions to the subject matter of this book come from selected
publications in diverse technical journals as well as from presentations at interna-
tional congresses on desalination and membrane technology held under the aegis of
the International Desalination Association (IDA) and the European Desalination
Society (EDS) as listed in the references for each chapter of the book.
The technical know-how that is brought together and described in this book thus
comes from a wealth of resources, while many experts contribute to the development
of this range of expertise as needed for the construction of reverse osmosis seawater
desalination plants. For this reason, only a general word of thanks can be addressed
to the many experts from the various disciplines who are involved in planning,
constructing, and optimizing the efficiency of these plants, which are of great
consequence for the water supply of humankind.
I am particularly grateful to Fichtner’s management for their technical and
financial support and for giving me access to the company’s many resources during
the preparation of this book and to the staff of the Fichtner Seawater Desalination
Division for the manifold exchanges of experience and, together with the other
divisions of the company, for the fruitful cooperation while tackling a large number
of projects. My thanks also go to Peter Billard for translating the German parts of the
manuscript into English and for his critical review of my English texts.
Last but not least, I would like to especially thank my wife for her tolerance and
patience during the time I worked on the book and for the support she gave the
family and me during my active professional life which involved much travelling
and frequent, sometimes prolonged, absences.

Nufringen, Germany Heinz Ludwig


May 2021
Contents

1 Introduction and Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


2 Earth’s Freshwater Resources and Their Management and the Role
of Seawater Desalination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1 Earth’s Water and Global Freshwater Distribution . . . . . . . . . . . . . 13
2.1.1 Global Water Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Distribution of Freshwater on Earth . . . . . . . . . . . . . . . . . 15
2.1.3 Global Warming and Its Influence on the Global Water
Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Global Withdrawal and Consumption of Freshwater and Water
Shortage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2.1 Global Freshwater Withdrawal and Water Consumption . . . 20
2.2.2 Water Shortage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3 Sustainable Water Management Systems and Their Components . . 29
2.3.1 Supply Management . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3.2 Water Demand Management . . . . . . . . . . . . . . . . . . . . . . 30
2.3.3 Integrated and Sustainable Water Management . . . . . . . . . 31
2.3.3.1 Wastewater Reuse . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.3.2 Seawater Desalination . . . . . . . . . . . . . . . . . . . . 36
2.3.3.3 Freshwater Resources in Integrated Water
Management and Their Allocation . . . . . . . . . . . . 37
2.4 Seawater Desalination Processes and Their Development . . . . . . . 45
2.4.1 Types of Desalination Processes and Their
Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.4.1.1 Thermal Processes . . . . . . . . . . . . . . . . . . . . . . . 45
2.4.1.2 Membrane Processes . . . . . . . . . . . . . . . . . . . . . 49
2.4.2 Development of Reverse Osmosis as a Seawater
Desalination Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.4.3 Potential and Future Prospects for Reverse Osmosis
in Seawater Desalination . . . . . . . . . . . . . . . . . . . . . . . . . 68
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

ix
x Contents

3 Seawater: Composition and Properties . . . . . . . . . . . . . . . . . . . . . . . 73


3.1 Composition and Temperature of Seawater . . . . . . . . . . . . . . . . . . 73
3.1.1 Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.1.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.2 Properties of Seawater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2.1 Thermophysical Properties . . . . . . . . . . . . . . . . . . . . . . . . 97
3.2.1.1 Specific Heat Capacity . . . . . . . . . . . . . . . . . . . . 97
3.2.1.2 Specific Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . 98
3.2.1.3 Thermal Conductivity . . . . . . . . . . . . . . . . . . . . . 99
3.2.1.4 Vapour Pressure . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.2.1.5 Boiling Point Elevation . . . . . . . . . . . . . . . . . . . 100
3.2.1.6 Latent Heat of Vaporization . . . . . . . . . . . . . . . . 102
3.2.2 Physical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.2.2.1 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.2.2.2 Dynamic Viscosity . . . . . . . . . . . . . . . . . . . . . . . 104
3.2.2.3 Kinematic Viscosity . . . . . . . . . . . . . . . . . . . . . . 104
3.2.2.4 Surface Tension . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.2.2.5 Solubility of Gases in Seawater . . . . . . . . . . . . . . 106
3.2.3 Physico-Chemical Properties . . . . . . . . . . . . . . . . . . . . . . 113
3.2.3.1 Ionic Strength . . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.2.3.2 Activity Coefficients and Solubility . . . . . . . . . . . 123
3.2.3.3 Osmotic Pressure and Osmotic Coefficient . . . . . . 148
3.2.4 Chemical Equilibria in Seawater . . . . . . . . . . . . . . . . . . . . 159
3.2.4.1 CO2/Bicarbonate/Carbonate Equilibrium . . . . . . . 159
3.2.4.2 Boric Acid/Borate Equilibrium . . . . . . . . . . . . . . 169
3.2.4.3 pH in Seawater . . . . . . . . . . . . . . . . . . . . . . . . . 172
3.2.4.4 Calcium Carbonate/Carbonate/Bicarbonate/CO2
Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Annexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
4 Seawater Reverse Osmosis (SWRO) Plant: General System
Configuration, Basic Design Parameters and Conditions, and
Overall Planning Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
4.1 SWRO General Systems Configuration . . . . . . . . . . . . . . . . . . . . 205
4.2 Basic Design Parameters and Conditions . . . . . . . . . . . . . . . . . . . 209
4.2.1 Basic Design Parameters, Their Acquisition and
Determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
4.2.1.1 Range of Composition and Quality of Seawater . . 213
4.2.1.2 SWRO Plant and Plant Systems Capacity and
Capacity Range . . . . . . . . . . . . . . . . . . . . . . . . . 227
4.2.1.3 Product Recovery Rate and RO Membrane
System Basic Equations . . . . . . . . . . . . . . . . . . . 242
4.2.1.4 Product Water Quality, Post-desalination,
and Post-treatment . . . . . . . . . . . . . . . . . . . . . . . 246
Contents xi

4.2.2 Basic Design Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 255


4.2.2.1 Overall SWRO Plant Flow and Mass Balances . . . 255
4.2.3 Basic Planning Requirements, Their Assessment and
Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
4.2.3.1 Energy Consumption and Supply Conditions . . . . 265
4.2.3.2 Chemicals Supply and Storage . . . . . . . . . . . . . . 270
4.2.3.3 Mode of Operation and Flexibility of the Plant . . . 273
4.2.3.4 Product Water Storage and Supply . . . . . . . . . . . 278
4.2.3.5 Environmental and Socioeconomic Impacts
and Their Influence on the Planning and Design
Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
4.3 Overall Planning Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
4.3.1 Phases of Planning and Design and Focus of Works
During the Planning Sequences . . . . . . . . . . . . . . . . . . . . . 307
Annexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
5 Reverse Osmosis Membrane System: Core Process of SWRO . . . . . . 315
5.1 Reverse Osmosis Membranes and Fundamentals of Their
Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
5.1.1 RO Membranes: Structure and Materials . . . . . . . . . . . . . . 317
5.1.1.1 RO Seawater Membrane Development History . . . 317
5.1.1.2 Membrane Structure and Material . . . . . . . . . . . . 321
5.1.2 RO Membrane Elements and Modules in Application
in SWRO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
5.1.2.1 Spiral-Wound Membrane Elements . . . . . . . . . . . 325
5.1.2.2 Hollow-Fibre Membrane Systems . . . . . . . . . . . . 330
5.1.3 Chemical and Mechanical Durability of Membrane
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
5.1.3.1 Mechanical Durability . . . . . . . . . . . . . . . . . . . . 333
5.1.3.2 Chemical Durability . . . . . . . . . . . . . . . . . . . . . . 345
5.1.4 Range of Operation Parameters and Conditions
of Polyamide and Cellulose Acetate Membranes . . . . . . . . 361
5.1.5 RO Membrane Design Basics . . . . . . . . . . . . . . . . . . . . . . 363
5.1.5.1 Water and Salt Transport Through RO
Membranes: Basic Equations . . . . . . . . . . . . . . . 364
5.1.5.2 Membrane Element Calculations . . . . . . . . . . . . . 381
5.2 RO Membrane Process Design . . . . . . . . . . . . . . . . . . . . . . . . . . 427
5.2.1 RO Membrane Module and RO System Configuration . . . . 427
5.2.2 Design Phases of an RO System . . . . . . . . . . . . . . . . . . . . 431
5.2.2.1 Phase 1: Definition of Design Conditions
and Basic Design Parameters . . . . . . . . . . . . . . . 432
xii Contents

5.2.2.2Phase 2: Calculation of the Framework


Parameters of the Membrane System
and of Approximate Values as Inputs for
Optimization and Detailed Calculation
of the Baseline System Design . . . . . . . . . . . . . . 436
5.2.2.3 Phase 3: Optimization of the RO System
Configuration, Detail Calculation of the
Performance of the Components and the
Composition of the Product Generated by
Membrane Elements, Modules, Stages,
and Arrays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
5.2.2.4 Phase 4: Extension of the Basic Design to Cover
the Full Ranges of TDS, Composition,
and Temperature of the Feed . . . . . . . . . . . . . . . . 454
5.2.3 Options for More Extensive Module, Array, and RO
System Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
5.2.3.1 Membrane Module Options . . . . . . . . . . . . . . . . 460
5.2.3.2 Array Options . . . . . . . . . . . . . . . . . . . . . . . . . . 465
5.2.3.3 System Options . . . . . . . . . . . . . . . . . . . . . . . . . 480
5.3 Fouling and Scaling in RO Systems . . . . . . . . . . . . . . . . . . . . . . . 485
5.3.1 Types of Fouling and Scaling with Their Influencing
Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486
5.3.2 Mechanism of Particle Transport, Deposition of Particles
on Membranes, and Properties of the Deposits . . . . . . . . . . 487
5.3.2.1 Particle Transport and Deposition . . . . . . . . . . . . 487
5.3.2.2 Properties of the Deposits . . . . . . . . . . . . . . . . . . 493
5.3.3 Fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
5.3.3.1 Fouling Potential of Raw and Pretreated Water:
Testing Methods . . . . . . . . . . . . . . . . . . . . . . . . 502
5.3.3.2 Membrane Performance Normalization and
Determination of Fouling Rate . . . . . . . . . . . . . . 518
5.3.3.3 Allowances for Fouling in Membrane Design . . . 529
5.3.4 Scaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
5.3.4.1 Scaling: Calculation and Prediction . . . . . . . . . . . 533
5.3.4.2 Scaling Control and Prevention . . . . . . . . . . . . . . 568
5.3.4.3 Uncertainties in Scaling Prediction Modelling . . . 588
5.4 RO Membrane Process Design with Membrane Manufacturers’
Design Software . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 596
5.4.1 Structure of RO Design Programs of the Membrane
Suppliers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 597
5.4.1.1 Necessary Input Parameters . . . . . . . . . . . . . . . . 597
5.4.1.2 Adjustment of Software Default and Estimated
Design Data and Parameters . . . . . . . . . . . . . . . . 599
Contents xiii

5.4.1.3 Results and Output Data of the Software


Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 600
5.4.2 Mode of Design Calculation and Application of the
Software for Changing Seawater Feed and Operation
Conditions of the SWRO . . . . . . . . . . . . . . . . . . . . . . . . . 603
5.4.3 Membrane Design Allowances and Safety Margins . . . . . . 604
5.4.3.1 Salt Rejection . . . . . . . . . . . . . . . . . . . . . . . . . . 604
5.4.3.2 Product Capacity . . . . . . . . . . . . . . . . . . . . . . . . 607
5.5 RO Process Units, Components and Overall System
Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
5.5.1 RO Process Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
5.5.2 RO Process Components . . . . . . . . . . . . . . . . . . . . . . . . . 611
5.5.2.1 RO Module Assembling Facilities . . . . . . . . . . . . 611
5.5.2.2 High Pressure Feed Supply and Energy Recovery
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
5.5.2.3 Safety Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
5.5.2.4 Chemicals Dosing . . . . . . . . . . . . . . . . . . . . . . . 676
5.5.2.5 Membrane Flushing, Cleaning,
and Preservation . . . . . . . . . . . . . . . . . . . . . . . . 690
5.5.2.6 RO Membrane Unit and System Configuration . . . 713
Annexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 721
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 732
Introduction and Overview
1

Water is vital to the development of life on Earth. The global evaporation-


precipitation water cycle transports fresh water from the oceans to the continents.
However, the global distribution of precipitation is very uneven and subject in many
regions of the world to significant dynamic and seasonal fluctuations. Ever since the
emergence of humankind, the availability and accessibility of fresh water have
shaped human settlement as well as agricultural and industrial exploitation of the
Earth’s continents. Until the recent past, the distribution of groundwater and surface
water and the frequency and intensity of precipitation were key factors for how
human settlement was structured as well as for economic and cultural development
throughout the world.
With the development of seawater desalination in recent decades, however, it has
become possible, at least in part and here in particular for arid areas as well as urban
regions close to the sea with a high population density, to sever the dependence of
fresh water supply via the global water cycle and to generate it directly from
seawater. Whereas in the arid regions of the Arabian Gulf, in the past this was
done mainly by thermal desalination processes coupled in dual-purpose plants with
electricity generation; more recently and extending up to the present, reverse osmo-
sis plants for seawater desalination, mostly constructed as single-purpose facilities
but also in hybrid dual-purpose configurations, are increasingly contributing to the
production of fresh water from the sea and are doing so worldwide. Seawater reverse
osmosis (SWRO) has now advanced to become the dominant technology for
obtaining fresh water from the sea worldwide, both in terms of the number of plants
installed and total desalinated water output.
The subject matter and scope of this book are the conceptual design and advanced
planning as well as the engineering of plants for the desalination of seawater by
applying reverse osmosis membrane technology together with the associated pro-
cesses for pretreatment of seawater and for post-treatment of the product water.
Students and teachers may find this manual informative and useful as an introduction
to reverse osmosis seawater desalination, but it is also intended as a working tool for
engineers, technicians, economists, and ecologists involved in the planning, design,

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 1


H. Ludwig, Reverse Osmosis Seawater Desalination Volume 1,
https://doi.org/10.1007/978-3-030-81931-6_1
2 1 Introduction and Overview

and operation of such desalination plants. It may also be consulted by the staff of
environmental agencies when considering issuing licenses for the erection and
operation of membrane desalination plants as well as by interested laypersons as a
source of information on the technical and ecological aspects of such plants.
The first part of Volume 1 of the book deals with strategic considerations
regarding water resource management in arid and urban regions and the role that
seawater desalination plays in these, followed by the physical and physico-chemical
properties of seawater that have to be known for the design and calculation of
seawater desalination plants. There follow descriptions of the planning and imple-
mentation of an SWRO project and the tasks of site selection and development of an
ecological concept for the facility that is appropriate for the location. The following
most comprehensive part of Volume 1 then describes the physical and physico-
chemical principles of reverse osmosis membrane calculations, the membrane
systems available for seawater desalination with reverse osmosis and their
characteristics, and the practical design of these systems as part of reverse osmosis
seawater desalination plants, including the associated techniques for energy recovery
and the necessary ancillary equipment.
In the following, an overview is presented of the subject matter of Chaps. 2, 3, 4,
and 5 of Volume 1 of this book.
Chapter 2 describes the global water cycle and how fresh water is distributed
around the world, as well as the likely impact of global warming on this cycle and the
consequential reduction in the availability of fresh water on the continents. In order
to provide sufficient fresh water as the global population rises even in arid and urban
regions, sustainable water management is called for, for the purposes of which
extraction from natural water resources and their supplementation through water
recycling and seawater desalination are appropriately controlled. The structure of
such integrated water management is described, and algorithms are presented that
can be used to determine the required share of seawater desalination for a sustainable
water supply in a given region. A further section presents a comparison of the
technologies for seawater desalination by thermal means versus reverse osmosis in
the form of process schematics, describes the respective treatment steps, and
compares their technical data and design parameters. For seawater desalination by
reverse osmosis, its development is then illustrated by noting and comparing the
growth in installed product water capacity of SWRO plants in the Arabian Gulf and
worldwide, with an appraisal of the potential for the future development of seawater
desalination by reverse osmosis, also under the aspect of coupling its energy supply
with regenerative energy facilities.
Chapter 3 begins with an outline of the range of seawater compositions based on a
number of analyses of seawater from various regions of the world (Annexes—
Annex 3.A1, Tables 3.15 and 3.16) followed by an explanation of how the aggregate
parameters of salinity (S) and total dissolved solids (TDS) for the salt content of
seawater are calculated. This is followed by quoting figures on the distribution of
temperature in various ocean regions around the world, the correlation between
seawater abstraction from different water depths and its temperature, and annual and
seasonal seawater temperature profiles at a range of desalination plant locations.
1 Introduction and Overview 3

When designing treatment processes for seawater desalination plants, the


seawater’s physical, thermophysical, and physico-chemical properties must be
known. All parameters in these categories of properties depend on salinity and
temperature. The equations used to calculate the values of the physical and
thermophysical parameters as functions of the salinity and temperature of the
seawater are given in the next two sections of Chap. 3. How these parameters vary
with salinity and temperature as calculated using the equations is depicted in graphs,
and the calculated values are additionally listed in tables in Annex 3.A2 of Chap. 3.
Described in the section dealing with the physico-chemical properties of seawater
are the basics for calculating solution equilibria and the solubility limits of the
components of high-salinity solutions by means of thermodynamic and stoichiomet-
ric modelling, as applied to the design of reverse osmosis plants and their
pretreatment processes. When calculating the physico-chemical properties and
how these vary with seawater salinity, the ionic strength I is also determined in
addition to the salinity S. How these are calculated from the molar or molal
concentration of the seawater components is shown at the beginning of this section.
Due to the high salinity of seawater and of the concentrates produced during
desalination, their physico-chemical properties differ greatly from those of fresh
water. The constituents of such high-salinity solutions are no longer present
completely in ionic form, but ion pairs are formed that influence how the seawater
components behave in solution. The fundamentals of ion association theory are
explained and their influence on the calculation of ionic strength is described.
For thermodynamic modelling of solution equilibria, it is necessary to know the
activity coefficients of the solution components as well as their corresponding
thermodynamic solubility products and how these vary with temperature. Calcula-
tion of the activity coefficients is done using the Debye-Hückel equation and its
derivatives. For high-salinity solutions, the specific ion-interaction algorithms are
used, these also being derived from the Debye-Hückel equation but are expanded to
take into account the more extensive ion interactions in such solutions. This chapter
includes the equations for thermodynamic modelling as well as a description of their
application ranges for different salinities. The Pitzer equations provide the algorithm
within the domain of specific ion-interaction theory that can be used to calculate the
activity coefficients of components even in highly saline solutions, such as reverse
osmosis concentrates. This system of equations is therefore used preferentially for
thermodynamic modelling of seawater and RO concentrates, so it is explained in
more detail in this chapter, which also includes a description of its application with
the aid of the PHREEQC calculation software. For calculating the thermodynamic
solubility product, polynomial equations are available for specific compounds. Such
equations are included in Chap. 3 for scalants such as barium sulphate BaSO4,
calcium sulphate CaSO4  2 H2O, strontium sulphate SrSO4, and calcium fluoride
CaF2. However, the solubility product can also be calculated from the standard
Gibbs free energy of reaction, and this in turn can be determined from the free energy
of formation of the reacting products and the reactants. These thermodynamic data
are compiled for each ion and compound in seawater, and this way of calculating the
solubility product is demonstrated taking strontium sulphate SrSO4 as an example.
4 1 Introduction and Overview

For stoichiometric modelling of the solubilities of seawater components and with


determination of their upper solubility limits, the respective stoichiometric solubility
product and its dependence on the salinity or ionic strength of the seawater as well as
on temperature and pressure must be known. Here, too, polynomial equations are
available for certain compounds with which the value of their solubility products can
be calculated as a function of the aforementioned parameters. This chapter provides
an equation for this purpose with differing coefficients for barium sulphate BaSO4,
strontium sulphate SrSO4, and a mixture of the two and shows how the value of the
solubility product varies with seawater salinity and temperature.
Osmotic pressure is a critical parameter for dimensioning RO installations. A
measure of how far it deviates from its behaviour in an ideal solution is the osmotic
coefficient. The calculation equations for osmotic pressure and the osmotic coeffi-
cient are stated and explained in detail in a dedicated section of Chap. 3. The
algorithms for determining the osmotic coefficient are derived in part from the
equations set up as a result of specific interaction theory. The various ways of
calculating this coefficient are presented, and their results are compared in a graph
with regard to their respective dependencies on salinity. The osmotic coefficient as
calculated using the Bromley polynomial is graphed as a function of temperature and
salinity and additionally listed in Annex 3.A2, Table 3.30 of Chap. 3. In another
graph, the osmotic pressure of seawater is plotted for the salinity range of 10–100 g/
kg and as a function of temperature from 10  C to 70  C.
Another section of Chap. 3 deals with the calculation of chemical equilibria in
seawater, in particular the CO2/bicarbonate/carbonate equilibrium, as well as the
boric acid/borate equilibrium. The equilibrium ratios in both cases are expressed and
calculated by stoichiometric equilibrium constants. For the CO2/bicarbonate/carbon-
ate equilibrium, these are the first and second stoichiometric dissociation constants
of carbonic acid and the stoichiometric ion product of water, whereas for the boric
acid/borate equilibrium, they are the stoichiometric dissociation constant of boric
acid. For both equilibria, the polynomials for calculating the stoichiometric constants
and their dependence on salinity and temperature are listed, how they vary with
salinity and temperature is depicted in graphs, and the calculated values are compiled
in tables in Annex 3.A2 of Chap. 3. Other graphs in this chapter show the depen-
dence of the proportions of the components of the two equilibria on pH and
temperature, and calculation equations are shown for determining the respective
concentrations of the components reacting in the equilibria. The concentrations of
the equilibrium reactants also influence the seawater’s pH, and how this is calculated
is described in another section of this chapter.
Calcium carbonate is a sparingly soluble component of seawater salinity, and, as
it is a scalant, it can adversely affect the desalination process in RO plants. Calcula-
tion of the calcium carbonate/carbonate/bicarbonate/CO2 equilibrium therefore
plays an important part in the design and operation of these installations. The
solubility of the crystalline manifestations of calcium carbonate in this equilibrium
is fixed by their thermodynamic or stoichiometric solubility product and the
concentrations of bicarbonate and carbonate. Both types of polynomial
equations—thermodynamic and stoichiometric—for calculating the solubility
1 Introduction and Overview 5

products of the various forms of calcium carbonate are given in the relevant section
of this chapter. In addition, calculation of the saturation pH and the saturation index
SI, which characterize the location of the equilibrium point towards solution or
precipitation of calcium carbonate, is shown using both thermodynamic and stoi-
chiometric modelling. Another measure of the tendency of calcium carbonate to go
into solution is the Stiff and Davis stability index, the calculation of which is also
presented.
Chapter 4 describes the planning and design activities during the development
phase of an SWRO project, namely, how the work proceeds in the course of its
conceptual and strategic planning. This planning phase starts with the determination
of the basic design parameters of the plant such as its net product water output
capacity and the range in which this is to be provided, the composition and quality of
the seawater to be desalinated, the recovery rate of the plant and of its reverse
osmosis tract, as well as the composition and quality of the product water. How the
required data on the seawater’s composition and quality can be obtained from
existing data sources and through a sampling and analysis campaign is described
in the first section of Chap. 4. Available analytical standards are listed and explained
with regard to their application for seawater analysis to determine its inorganic
components and to analyse its other quality-critical constituents. This section also
describes pilot plant trials of various potential processes for seawater pretreatment
and the evaluation of the information thus obtained to select the optimum combina-
tion of processes for this SWRO stage.
For the conceptual design of the plant’s treatment stages, the availability on the
component and SWRO process levels as well as, as a consequence, on the availabil-
ity of the plant as a whole is also critical, and their influence on plant design is
described in a further section. There it is also explained how the range of product
water output flow at which the desalination plant is to be operated influences its
technical design, especially if its energy supply is provided by renewable energy
generation without backup from fossil energy.
Another section of Chap. 4 contains the basic equations for the RO processes,
with calculations of the recovery rate, the concentration factor, and the dimensioning
of a second post-desalination pass, with calculation of the capacity factor for this RO
stage and the impact of the recirculation of concentrate from this pass into the
seawater RO feed line on the recovery rate of the RO tract and of the entire plant.
Likewise influencing the recovery rate of an SWRO plant is the recovery rate of the
pretreatment stage, as also shown in this section.
As conceptual planning nears its conclusion, the first process flow diagrams of the
entire plant are prepared, showing flow and material balances. The corresponding
systems of equations for calculating the outputs and concentrations in these diagrams
are brought together in another section, in which the flow and concentration diagram
of an SWRO plant with a product water output of 100,000 m3/day is presented by
way of example. This 100,000 m3/day SWRO plant is also used as a reference in
subsequent chapters to provide examples of the aspects of SWRO plant design that
are the subject of these chapters.
6 1 Introduction and Overview

A further essential component of the basic engineering of an SWRO plant is the


determination of its energy source, namely, whether it is to be connected to an
existing fossil-based power network that supplies its entire needs, whether this
network is merely to serve as a transmission or backup system for the plant’s
regenerative energy supply, or whether the desalination plant is to be supplied
entirely from regenerative energy sources.
Further conceptual aspects are:

• The configuration of the supply of chemicals and their storage and dosage
together with the provision of their in-plant water consumption,
• Storage of the generated product water and dimensioning of the storage capacity
under the aspects of security of supply of the water offtakers and availability of
water for SWRO commissioning, to cover emergencies, and as fire-fighting
water.

Towards the end of the conceptual and strategic planning phase of the SWRO
plant, sufficient information about its configuration, design, and process flows is
available to decide on its location. The plant can be integrated into existing indus-
trial, tourist, and commercial complexes or into a power plant infrastructure, or it can
be designed as a self-sufficient standalone facility. High-capacity municipal SWRO
plants for drinking water supply are often realized in the form of the latter option. If
there are several locations to choose from for such a project, the one that is most
suitable has to be identified through a structured comparison approach. Technical
and economic requirements as well as environmental protection, public health
issues, socioeconomic aspects, and project risks must be taken into account. The
principal and subsidiary criteria are listed in another section of this chapter, and the
methodology for such a complex comparison takes the form of multi-criteria analy-
sis (MCA) by applying the weighted sum method (WSM).
Another section of Chap. 4 describes the potential environmental and socioeco-
nomic impacts of an SWRO plant on its surroundings. Based on these findings and
together with the conceptual and strategic planning, an environmental action plan is
drawn up with the aim of largely mitigating adverse impacts or eliminating them
altogether by taking suitable measures when designing and configuring the plant.
This planning provides the framework for the environmental licensing procedure for
construction and operation of the desalination plant. An important tool for structur-
ing and organizing permit planning and for compiling the documentation for the
permit application procedure is the environmental impact assessment (EIA), whose
content and workflow for its preparation are explained in another section of this
chapter and also presented as a schematic.
Chapter 4 concludes with a general overview of the process for planning an
SWRO plant and a description of the planning workflow and its design phases as
well as of the outcomes and documentation that results from the various project
phases. The sequence of project phases and the activities to be carried out in each
with the resulting documentation are shown in a schematic diagram.
1 Introduction and Overview 7

Chapter 5 which is the most extensive chapter of this book, starts with a
description of osmotic processes in a system of two solutions of differing
concentrations separated by a semipermeable membrane and the reversal of this
process for its application as reverse osmosis (RO) for the desalination of solutions.
A historical overview describes membrane development up to the advent of reverse
osmosis membrane elements for brackish water and seawater desalination. The RO
membranes used today for this purpose take the form of thin-layer polyamide
composite membranes and cellulose acetate hollow-fibre membranes. Their applica-
tion in practice as spiral-wound elements, in particular polyamide composite
membranes or cellulose acetate membranes that are aggregated as hollow-fibre
bundle elements, and the designs of these are described. A number of these mem-
brane elements are housed in a pressure vessel that is termed a membrane module in
which, on the concentrate side, they are connected in series. The RO desalination
units are then made up of the membrane modules, which are connected in parallel
and sometimes in a stepwise arrangement.
The mechanical stability of the membrane modules and thus the pressure to which
they can be subjected depend on the physical structure of the separation membranes
within the elements, the construction of the membrane elements themselves and the
pressure they can withstand, and the strength of the pressure vessel housing. When
designing an RO stage, safeguards must also be incorporated to prevent mechanical
damage to its membranes, for example, due to telescoping, pressure surges caused
by insufficient venting, and excessive backpressure on the permeate side.
The chemical resistance of the membrane elements and membrane modules is
primarily determined by the membrane materials but is also influenced by the
materials used for the element structure and for the module housing. Chapter 5
describes in particular the chemical interactions of the oxidizing disinfectants,
chlorine, chlorine dioxide, and monochloramine, that are used for disinfecting
seawater; their influence on the membrane materials polyamide and cellulose ace-
tate; and the difference in the resistance of the two materials to these agents. The
resistance over the acidic to the alkaline pH range of the two membrane materials is
also compared. Likewise having a significant influence on the resistance of the
membrane materials is the temperature during chemical treatment. The membrane
manufacturers specify permissible physical and chemical conditions for their mem-
brane elements, and these are tabulated for the polyamide thin-film composite (TFC)
membranes and the hollow fine fibre (HFF) cellulose triacetate membranes.
The subsequent sections of Chap. 5 describe the design of the RO assemblies of
an SWRO plant, starting with the calculation fundamentals of RO membranes to set
out the theoretical principles and the calculation models for water and salt transport
through RO membranes and how these two parameters are influenced by the salinity
and temperature of the water to be treated, the operational desalination recovery rate,
the pressure applied to the membrane, and the pressure and salinity of its permeate.
The concentration polarization on the membrane feed side is explained and how this
factor is calculated is shown. It has a significant influence on the salt concentration
on the membrane wall and thus on the permeability and salt rejection of an RO
membrane.
8 1 Introduction and Overview

When calculating a membrane element, the point-by-point, one-dimensional


theoretical calculation models for RO membranes have to be transformed into a
two-dimensional calculation, i.e., the water and salt fluxes and the element’s con-
centration polarization have to be calculated from the variation of the feed/concen-
trate concentrations and the concentration ratios across the element’s surface.
Further, the calculation has to be matched to the conditions encountered in practice
during the operation of a membrane element, for example, the pressure loss in the
element’s flow channels and the reduction of the permeability and salt rejection of
RO membranes due to ageing have to be taken into account. The calculation
algorithms for this can be found in the relevant sections of Chap. 5.
The product flow and salt rejection of commercial RO membrane elements are
specified by their manufacturers on the basis of “standard test conditions” with
defined values for salinity expressed as mg/l NaCl, pressure, temperature, and
element recovery rate. From these standard values, the element’s performance
figures under design conditions of salinity, pressure, temperature, and membrane
operating time are determined with the above-mentioned calculation equations,
although these figures are adjusted accordingly by applying correction factors for
the influences of temperature and membrane age. The procedure for calculating a
membrane element’s performance parameters for product flux and product compo-
sition under specified design and operating conditions over a defined membrane
element operating time from its standard parameters as stated by the membrane
manufacturer is tabulated as a flow chart.
There are additional algorithms shown for calculating the salt passage of the
components of the CO2/bicarbonate/carbonate equilibrium and the boric acid/borate
equilibrium with their dependence on pH.
For the design of an RO installation, guidelines and limits are prescribed by the
membrane manufacturers for the use of their membrane elements, stating the maxi-
mum values of various design and operating parameters such as average product
flux, element recovery rate, feed and concentrate flows, and the concentration
polarization factor. These values are specified to avoid increased concentration
polarization in the elements, to ensure their uniform flow, and to prevent pressure
loss across the elements from exceeding a maximum value, which is also specified.
For the seawater membrane elements in the first RO pass, the value ranges of these
guidelines are oriented to the type of seawater extraction and its pretreatment and, for
the brackish water membrane elements for post-desalination in the second pass, to
the admission of first-pass product water. These membrane manufacturers’ design
guidelines are compiled in a table.
When calculating an RO configuration, the number of membrane elements with
their arrangement in the membrane modules and that of the modules within the RO
installation must be such that these maximum guideline values will not be exceeded
and that the RO operating values will remain within the specified range of these
parameters. The calculations required to achieve this must be done iteratively, with
the performance data for the RO installation regarding product flow, recovery rate,
and composition being specified, and the design process commencing with approxi-
mate values for the number of elements in the plant and in the modules as well as for
1 Introduction and Overview 9

the RO operating pressure, and then adjusting these parameters in a number of


recalculation steps until the specified design criteria are met. The membrane
modules’ product water flow and composition are calculated from the flows and
compositions of the part streams from the module elements, while the performance
figures for the RO installation are calculated from the part-stream flows from the
membrane modules, with the installation’s design adjusted to match its specified
product performance. The algorithms required to do this and the calculation
procedures are stated and explained in the relevant section of this book. Because
these complex design calculations are so laborious and time-consuming that it is no
longer practical to perform them manually to the level of detail required for the
design of an RO installation, the membrane manufacturers provide calculation
software for this purpose. Their basic structure with the necessary input data and
their output are explained and shown schematically, and how these software tools are
used as design aids is described.
There are several possibilities for optimizing a reverse osmosis design with regard
to process performance and energy consumption. At the level of the membrane
modules, this is firstly the choice of the number of membrane elements with which
the pressure vessels are equipped and, secondly, installing elements with differing
salt rejections in the front and in the rear of the modules, termed hybrid element
internal staging. At the levels of RO arrays and the plant, this is the split partial mode
to reduce the post-desalination capacity that has to be installed and product throt-
tling, which is the build-up of product-side backpressure during RO operation to
simplify control of the feed-side operating pressure. Further, the seawater desalina-
tion pass of an SWRO plant can be configured in two stages with the second stage
equipped with special seawater RO membranes that can withstand a higher pressure.
This stage is operated with an additional booster pump, thus achieving a higher
recovery rate. Optimization at plant level is also possible by adjusting the frequency
of membrane replacement and thus the membrane age to so influence RO membrane
permeability and salt rejection. These optimization options are described in detail in
Chap. 5 with regard to their influence on plant design and their efficacy.
Due to fouling and/or scaling during RO operation, deposits are formed on the
membrane surfaces which negatively impact membrane permeability and salt rejec-
tion. The fundamentals of the transport of particles and their deposition on mem-
brane walls and in flow channels, the flow resistance of such layers and their
influence on the concentration polarization of ionogenic components, and the
associated effect on membrane water and salt transport, termed cake-enhanced
osmotic pressure, are dealt with in a further section of Chap. 5. Since changes in
temperature and salinity as well as RO membrane age also impact the flow and
composition of product water from an RO tract, in order to quantify fouling-induced
changes in membrane performance, the actual operating and performance data of an
RO installation have to be normalized to standard or reference values for salinity,
temperature, and membrane age using normalization algorithms. If after this nor-
malization RO performance deviates from the design performance specifications for
the standard or reference conditions, the occurrence of membrane fouling has to be
assumed. The calculation algorithms for this data normalization of an RO installation
10 1 Introduction and Overview

are included in the relevant section of Chap. 5. If, after chemical cleaning of the RO
membranes, the design data are no longer achieved under standard conditions, this
residual fouling is referred to as “irreversible fouling,” and the resulting reduction in
performance must be additionally considered when dimensioning an RO installation.
To what extent and how this should be taken into account for the design of an RO
installation when determining its maximum inlet pressure and/or its membrane area
are also explained in Chap. 5.
Prevention of scaling due to the concentrates generated during desalination is
crucial for reliable operation of an RO installation. The scalants contained in
seawater are alkaline earth sulphates, calcium fluoride, and calcium carbonate. The
scaling potentials of these various scalants are characterized by the saturation ratio
SR, the saturation index SI, or the supersaturation ratio SSR, and these depend on the
solubility product of the respective scalant and on how its concentration increases
during desalination. The value of the solubility product of a scalant that becomes
established in the RO concentrate depends on the concentrate’s salinity and temper-
ature. In the section of Chap. 5 that deals with scaling, it is shown how the values of
these parameters are calculated for the alkaline earth sulphate scalants, for calcium
fluoride, for calcium carbonate, and also for magnesium hydroxide and for a
low-solubility mixture of magnesium compounds in the alkaline operating mode
of the second pass by means of both stoichiometric and thermodynamic modelling.
For stoichiometric modelling, algorithms for the calculation of the stoichiometric
solubility product as a function of salinity and temperature are included in this
section for the above compounds and additionally as a function of pH for magnesium
hydroxide. For calculating the acid dosage into the RO feed line to suppress calcium
carbonate scaling, a target value for the Stiff and Davis stability index (S&DSI) in
the concentrate is specified. This section includes a table of the steps together with
the respective algorithms for calculating this parameter, which is a measure of the
scaling potential of calcium carbonate in the RO concentrate. The acid dosing rate
that is needed to achieve the S&DSI target value in the RO concentrate must be
calculated iteratively, as shown in a flow chart.
For thermodynamic modelling of scaling and calculation of the thermodynamic
saturation index SITh for calcium carbonate scaling as well as for the alkaline earth
scalants, the activity coefficients of the components involved must be known, these
being determined by applying the specific interaction algorithms of the Pitzer
equations that take into account the high salinity of the RO concentrate. The
PHREEQC software tool with the data package pitzer.dat is used for the complex
calculations involved in this approach to modelling the scaling process.
The formation of scale by the alkaline earth sulphates and calcium fluoride in the
first pass and by magnesium compounds in the second pass of an SWRO plant is
suppressed by dosing antiscalants into the RO feed line. Calcium carbonate scaling
can also be inhibited by dosing antiscalants, and this is applied for the alkaline mode
of SWRO operation, for which only antiscalant is dosed. Antiscalants serve to retard
the precipitation of scaling products from supersaturated solutions. The kinetics of
these processes, the calculation of the dosing concentration of the antiscalants, and
the supersaturation in the RO concentrate as allowed by the membrane
1 Introduction and Overview 11

manufacturers when dosing these products are listed and explained in the relevant
section of Chap. 5.
The process-oriented section of Chap. 5 describes the design of the RO tract and
the types and selection of the process-engineering components with which the
membrane installations are operated.
The pump types, their energy efficiency and the manner of pressure and power
control of the pumping stations selected for the RO feed, and the selection of the
energy recovery devices for the RO seawater desalination stage and their efficiency
are decisive factors for an SWRO plant’s energy consumption. For the high-pressure
centrifugal pumps predominantly used for feeding medium- and high-capacity
seawater desalination installations, the basics of pump design and, in particular,
the setting of delivery and operating pressure by means of variable speed control are
described. Further, the potential efficiency of the pumping units and its dependence
on pump design and manner of operation are explained.
The technologies of the three most common energy recovery devices (ERDs):

• Pelton turbine
• The isobaric ERDs work exchanger and pressure exchanger
• Turbochargers

These technologies are described and how they work is explained. The algorithms
for calculating the SWRO plant’s energy consumption with each ERD type are
presented.
This section of Chap. 5 also describes the design of the dosing stations for dosing
chlorine and then bisulphite for subsequent dechlorination, with selection of the
dosing points, the chemical demand at the dosing points, and the respective dosing
rates. Also shown are the equations for calculating the energy demand of each dosing
station of the RO tract and the total energy consumption of all RO dosing stations.
Chemical cleaning of the RO membranes is necessary to remove reversible
fouling from their surfaces. The chemical cleaning solutions are selected depending
on the nature and composition of the fouling deposits. In a table in the respective
section of Chap. 5, formulations recommended by membrane manufacturers for the
removal of certain types of fouling are compiled, stating the operating conditions
regarding pH and temperature to be maintained during the cleaning process. Further,
this section also contains the basic data and calculation equations for dimensioning
the cleaning stations and a description of the cleaning processes with the required
times and amounts of wastewater generated during cleaning. Also found in this
section are the calculations of the energy demand for chemical cleaning and a
description of measures that can be taken to raise cleaning efficiency, namely, gas
bubble support during cleaning, direct osmosis-osmotic back flushing, and reverse
cleaning. The chemical cleaning station is also used for mothballing an RO tract
during prolonged shutdowns. Conservation biocides together with the respective
conditions for plant preservation are compiled in a table.
Chapter 5 concludes with an overview of the possibilities for modular construc-
tion and array configurations of RO tracts, such as their configuration with multiple
12 1 Introduction and Overview

trains, in a feed-centre configuration, or in three-centre configuration. The pros and


cons of these different configurations for the design and operation of an SWRO plant
are explained and commented on.
Earth’s Freshwater Resources and Their
Management and the Role of Seawater 2
Desalination

2.1 Earth’s Water and Global Freshwater Distribution

Earth is a water planet, with 70% of its surface being covered by water (Fig. 2.1).
The saltwater of the oceans is home to a huge diversity of life. Yet also on the
Earth’s land masses, water is the basis of all plant, animal, and human life.
The mean value of salt content of seawater ranges from 7 g/l in the Baltic Sea to
between 30 and 40 g/l in the great oceans. At 45 g/l, salt content in the Red Sea and
Persian Gulf is even higher. However, life on Earth’s continents needs freshwater
with a substantially lower salt concentration. The most favourable mineral content in
freshwater varies according to genus and species. Table 2.1 presents an overview of
the tolerable salt content for various life forms on Earth’s continents. Of course, salt
content is not the only factor that determines the optimum compatibility of freshwa-
ter for the various species. A further role is played by the individual components of
the salt content.

2.1.1 Global Water Cycle [1]

The global water cycle transports the freshwater essential for life on Earth from the
oceans to the continents. Seawater evaporation, a natural process of distillation, is
the source of freshwater within this cycle. Thus, water in its three phases, vapour,

Table 2.1 Tolerable salt Range of tolerable salt


content of water for various Life forms content (mg/l)
life forms
Plants and crops 300–3000
Animals and livestock 2000–10,000
Freshwater fish 4000–12,000
Drinking water for humans 100–1000

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 13


H. Ludwig, Reverse Osmosis Seawater Desalination Volume 1,
https://doi.org/10.1007/978-3-030-81931-6_2
14 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.1 Earth—a watery planet (source: NASA- visibleearth.nasa.gov—The Blue Marble)

liquid, and ice, is the primary energy transport medium in the Earth’s energy
balance. As it changes between its three phases, water absorbs and transports energy
before releasing it. The water vapour and heat absorbed during seawater evaporation
are distributed across the continents by wind currents. Condensation of the water
vapour, accompanied by the formation of rain, snow, and ice, results in the gravity-
controlled return of water to the sea, where the cycle begins anew.
Precipitation falls on the land in the form of rain or snow. Some of it penetrates
into the subsoil, where it is stored as groundwater. Some of it collects in streams,
rivers, and lakes before flowing on into the sea. However, a large proportion of
precipitation is deposited as ice and snow in the Earth’s polar regions and mountains.
Also, some precipitation re-evaporates before or after it reaches the surface of the
planet and is transported further in the form of water vapour.
Figure 2.2 presents the processes of the global water cycle and indicates the
quantities of water transported in the different parts of the cycle.
2.1 Earth’s Water and Global Freshwater Distribution 15

Fig. 2.2 Global water cycle (source: UNEP/GRID-Arendal, Philippe Rekacewicz; Igor
A. Shiklomanov, State Hydrological Institute (SHI, St. Petersburg) and United Nations Educational,
Scientific and Cultural Organisation (UNESCO, Paris), 1999). Note: The width of the blue and grey
arrows is proportional to the volumes of transported water

Within the Earth’s global water cycle, nature not only transports freshwater from
the oceans to the land, but it also transports and distributes energy to all parts of the
planet. The water cycle, therefore, is primarily an energy transport mechanism by
which, on reaching the Earth, the sun’s energy is distributed around the planet.
However, it is also the case that any changes in the Earth’s energy balance will have
a significant effect on the planet’s water transport mechanisms and meteorological
conditions.

2.1.2 Distribution of Freshwater on Earth [2]

Of the water in the Earth’s oceans, 97% is saltwater, with only around 2.5% being
transported to the continents as freshwater by the global water cycle. It is estimated
that, of this quantity of freshwater, around 70% is in the form of ice at the north and
16 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.3 Global freshwater


and its distribution (data
source: UNESCO World
Water Development Report
WWDR 2 (2006), based on
data from Shiklomanov and
Rodda 2003)

south poles or as snow and glaciers on the Earth’s mountains, while over 30% is
available as groundwater and in aquifers. The remainder is found in lakes and rivers
or as moisture in soil, at the Earth’s surface, and in vegetation (Fig. 2.3).
If precipitation was uniformly distributed across the planet, the amount of water
transported by the global water cycle would be sufficient to supply all of the Earth’s
continents with enough freshwater. However, the global distribution of precipitation
is highly non-uniform and subject in many regions of the world to significant
dynamic and seasonal changes. Annual precipitation in the different climate zones
ranges from less than 100 mm in desert regions to over 2500 mm in subtropical
zones (Fig. 2.4).
2.1 Earth’s Water and Global Freshwater Distribution 17

Fig. 2.4 Distribution of mean precipitation on Earth (1961–1990) (source: Gridded Precipitation
Normals Data Set, Global Precipitation Climatology Centre (GPCC), Offenbach 2007)

Table 2.2 Aridity index and arid zones and drylands on Earth
Precipitation
Climate Aridity Annual Annual wet periods % of world % of total
zone index (AI) precipitation (months/year) land area drylands
Dry 0.50–0.65 600–1200 mm 4–6 ~10 21
sub-humid
Semi-arid 0.20–0.50 400–600 mm 3–4 ~18 37
Arid 0.05–0.20 100–400 mm 1–3 ~12 25
Hyper-arid <0.05 <100 mm 0–1 ~8 17
(desert)
Drylands <0.65 ~48 100
Arid zones <0.50 ~38 79

Non-uniform distribution of precipitation over the Earth’s continents has led to


vast areas of land with low rainfall (arid zones). Based on an aridity index (AI), these
areas are classified into the following climate zones (Table 2.2) [3–5]:

• Dry sub-humid
• Semi-arid
• Arid
• Hyper-arid

The aridity index (AI) is the quotient of mean annual precipitation P and mean
annual potential evapotranspiration PET (Eq. 2.1):
18 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

P
AI ¼ : ð2:1Þ
PET

AI ¼ aridity index
P ¼ mean annual precipitation
PET ¼ mean annual potential evapotranspiration

Evapotranspiration is the sum of mean annual transpiration, evaporation, and


vaporization of water from the surface of the soil as well as from flora and fauna in
the relevant climate zone. It is influenced mainly by the temperature conditions in the
respective arid zone as well as by the intensity and duration of solar radiation. In an
arid climate zone, annual potential evapotranspiration exceeds annual precipitation.
This means that when rain falls, most of the moisture evaporates and returned to the
atmosphere.
In such regions, rainfall is highly irregular with long intervals between precipita-
tion events. Consequently, agriculture in arid climate zones is possible, if at all, only
with irrigation. Also in semi-arid regions, farming is mainly dependent on irrigation
owing to the irregularity of rainfall.
Arid (arid and hyper-arid) and semi-arid climate zones together make up as much
as 38% of the Earth’s land area. If dry sub-humid zones, some of which are also of a
semi-arid character, are included, then drylands account for up to 48% of the Earth’s
land masses (Fig. 2.5).
Groundwater systems (aquifers and other near-surface groundwater storage)
contain over 30% of the globally available quantity of freshwater [6]. In terms of
freshwater that is available in liquid form, the groundwater fraction is even as high as
98%. This means that the available groundwater reserves are many times greater

Fig. 2.5 Global distribution of arid zones and drylands (source: World Soil Resources Map,
FAO/EC/ISRIC, 2003)
2.1 Earth’s Water and Global Freshwater Distribution 19

Fig. 2.6 Large aquifers of the world (source: Jean Margat, Great aquifer systems of the world. In
Aquifer Systems Management: Darcy’s Legacy in a World of Impending Water Shortage. Chery
Laurence and Ghislain de Marsily (eds.) Oxford, England: Taylor & Frances, 2007, p. 105)

than the quantity of surface water in lakes and rivers. Many aquifers are distributed
across the Earth’s continents (Fig. 2.6).

2.1.3 Global Warming and Its Influence on the Global Water Cycle
[7, 8]

It is by now undisputed that global warming is already leading and will continue in
future to lead to climatic changes on the continents of our planet. However, the
highly complex phenomena involved in meteorological processes and the associated
global water cycle make it extremely difficult to model and predict the nature or scale
of the changes that can be expected. Yet there is agreement among climate
researchers that global warming will have an influence on the global water cycle.
The likely consequences are:

• Changes in the frequency and global distribution of precipitation


• Increased uncertainty with regard to precipitation forecasting and also with regard
to determination of the availability of freshwater, particularly in semi-arid regions
of the Earth
• Increased winter runoffs in snow-rich regions together with reduced runoffs in
summer because of changes in the extent and timing of snowmelt and glacial
retreat in the mountains
• Reduction in the extent and frequency of precipitation in arid regions
• Extension of semi-arid zones into sub-humid regions and/or transformation of
semi-arid into arid climate zones as a result of reduced frequency and intensity of
precipitation and increased evapotranspiration
20 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

• Declining replenishment of groundwater reserves, especially in semi-arid regions,


coupled with reduced availability of freshwater resources.

Climate change prediction models point to trends towards increased precipitation


in tropical regions and reduced precipitation in drylands. Also, precipitation in both
of these climate zones will be subject in future to significantly greater variation.
Many predictions estimate that global warming will lead to an intensification,
acceleration, and increase of the global water cycle. There are already discernible
signs of this today, such as the growing number of extreme weather events around
the world, including in temperate climate zones. At the same time, however, global
warming will lead to increased evapotranspiration in the arid zones of the Earth.
Climate models predict that a higher evaporation rate in these zones will accelerate
the trend towards desertification.
Changes in the water cycle as well as in global and regional patterns of rainfall
distribution, especially in tropical regions, will have an influence on soil moisture
and there will be increased soil erosion.
It can generally be expected that global warming and climate change will affect
the patterns of global freshwater distribution to such an extent that especially the arid
regions of the Earth will suffer a decline in available freshwater resources.

2.2 Global Withdrawal and Consumption of Freshwater


and Water Shortage

2.2.1 Global Freshwater Withdrawal and Water Consumption


[8–10]

Globally, humans withdraw around 8% of the Earth’s renewable freshwater


resources for use in agriculture as well as for domestic and industrial consumption.
The only renewable freshwater is the one which the global water cycle transports
onto land, where it falls as precipitation. This quantity of freshwater is made up of
two parts. One part, having fallen as precipitation, is absorbed by the soil or by plants
and organisms and is then cycled back to the atmosphere (around 60% of freshwater
resources). This so-called Green Water is the main source of water supply to natural
ecosystems as well as for natural irrigation of agricultural land. The second part,
“Blue Water”, collects in surface watercourses, some of it serving also to replenish
the groundwater reserves (around 40% of renewable freshwater resources).
Although all of the water accessible to humans is withdrawn from “Blue Water”,
only around 30% of “Blue Water” is available to humans for withdrawal as
exploitable water. Globally, over 50% of this freshwater fraction is already
exploited.
Of all the freshwater withdrawn globally in 1995, around 66% was used in
agriculture (mainly for irrigation), 9% for industrial consumption, and 20% for
domestic needs (see Table 2.3).
2.2 Global Withdrawal and Consumption of Freshwater and Water Shortage 21

Table 2.3 Withdrawal and consumption rates of water users, 1995 (data source: UNESCO—
International Hydrological Program—World water resources and their joint use; I.A. Shiklomanov,
State Hydrological Institute (SHI), St. Petersburg, 1999)
Global average withdrawal rate of water Consumption rate of
Sector users (%) withdrawal (%)
Agriculture ~66 ~46
Industry ~9 ~2
Domestic ~20 ~2
Reservoir ~5 –
losses
Consumption rate of total withdrawal ~50

In agriculture, however, average actual consumption, i.e., the water actually used,
was around 55% of the amount withdrawn, while, for industry and households, the
figure was as low as around 10–15%. Agriculture and artificial irrigation, therefore,
are the main consumers of the Earth’s freshwater resources. If precipitation on
agricultural land from “Green Water” is added to the amount of “Blue Water”
used for artificial irrigation, then freshwater consumption for natural and artificial
irrigation accounts for over 80% of available freshwater resources [11].
The majority of water use for irrigation is consumptive, i.e., that water is not
available for reuse. Some of it evaporates into the atmosphere and is thereby returned
to the global water cycle, while another part of it is retained in the irrigated crops.
The rate of return flow of water from artificial irrigation systems to surface
watercourses is low. Also, the quality of such water is diminished. Due to high
evaporation rates in semi-arid and arid zones, artificial irrigation makes virtually no
contribution to groundwater replenishment in those regions.
Up until the middle of the last century, global freshwater withdrawal was more or
less in line with the Earth’s population growth. In the meantime, however, growth in
freshwater consumption by humans is increasingly exponential in comparison with
global population growth (Fig. 2.7).
This means that water withdrawal from the Earth’s renewable freshwater reserves
will rise to around 12% by 2025. Viewed in relation to the globally accessible
amount of freshwater so by that date, some 80% of those reserves will be withdrawn
by humans as exploitable water.
The situation is similar with regard to the share of withdrawal and consumption of
freshwater, respectively, for agriculture and for industrial and domestic use (see
Fig. 2.8).
The grey areas of the charts indicate the difference between water withdrawal and
actual consumption for the three user groups.
Agricultural consumption, especially consumption for artificial irrigation, has to
date been the key factor with regard to water withdrawal from accessible global
freshwater resources. To guarantee food supplies to the Earth’s growing human
population, the area of land under cultivation will continue to increase. There will be
a need to develop semi-arid, and possibly also arid, zones for agricultural use. Global
warming, too, may make it necessary for existing cultivated areas to be additionally
22 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.7 Global water withdrawal and population growth (data source: Global Water Intelligence
(GWI)/UN/FAO Aquastat)

irrigated by artificial means. Agriculture will, therefore, withdraw increasing


amounts of freshwater from the global water cycle. It is estimated that, by 2025,
the amount of freshwater withdrawn for agricultural use will increase by up to 1.3-
fold. Industrial freshwater use, which is significantly lower than for agriculture, is
predicted to grow by up to 1.8-fold, with domestic use set to rise by up to 1.5-fold.
However, the availability of water resources as well as the type and extent of
water use vary considerably from continent to continent and from region to region.
Consequently, globally based studies and predictive models can provide only a very
limited indication of actual regional and local water conditions. Figure 2.9 shows
agriculturally used water as a percentage of total freshwater withdrawal for various
countries and regions. Particularly in semi-arid and arid climate zones, the fraction of
agriculturally used freshwater sometimes makes up between 75% and 90% of
available water resources [12].
The same applies to densely populated regions of the world, particularly in Asia,
where water use by agriculture places a similarly significant stress on available water
reserves. In many cases, water withdrawal significantly exceeds the extent of local or
regional natural resources.
Where the availability of “Blue Water” from surface watercourses is no longer
sufficient to meet the needs of agriculture, industry, and municipalities, growing use
is being made of groundwater resources [13]. As a global average, up to 50% of
municipal water consumption, 40% of use by industry, and 20% of water use for
artificial irrigation are based on groundwater withdrawal. However, the percentage
2.2 Global Withdrawal and Consumption of Freshwater and Water Shortage 23

Fig. 2.8 Current and predicted global freshwater withdrawal by user group (source: Philippe
Rekacewicz, UNEP GRID ARENDAL; Igor A. Shiklomanow, State Hydrological Institute (SHI,
St. Petersburg) and United Nations Educational, Scientific and Cultural Organisation (UNESCO,
Paris), 1999)

use of groundwater varies considerably from region to region. Water in many cities
and megacities is supplied almost exclusively from groundwater. Also, a high
percentage of farmed land is irrigated solely with groundwater.

2.2.2 Water Shortage

Water is vital to the development of life on Earth. The existence, availability, and
accessibility of freshwater shape the patterns of human settlement as well as agricul-
tural and industrial use of the Earth’s continents. Up until recent times, the distribu-
tion of groundwater and surface water and the frequency and intensity of regional
precipitation were crucial with regard to not only the settlement structures but also
the economic and cultural development of humankind. Against this background,
regional water shortages and droughts occurred predominantly as a consequence of
natural seasonal and climatic variations in respect of the kind and extent of precipi-
tation. However, population growth, allied to the intensified agricultural use of land
and increased exploitation of mineral resources, has led to the creation of more and
more local centres of high population density. Human habitats have also expanded
24 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.9 Agriculturally used water as a percentage of total freshwater withdrawal, 2001 (source:
UNEP/GRID-Arendal, IAASTD/Ketill Berger; FAO, Aquastat, 2007)

into areas with fewer and non-renewable freshwater resources. This trend has
resulted more recently in water stress or water shortage in many regions of the
world. Water stress refers to an imbalance between the necessary withdrawal of
freshwater to meet a region’s water demand and the locally available water
resources. Water stress is also an indicator of the degree of strain on a region’s
water resources and water ecosystems from external factors.
A region’s water supply situation or level of water stress can be characterized by
water stress indices. One such index is the withdrawal-to-availability (WTA) ratio or
water stress indicator (WSI), i.e. the ratio of annual freshwater withdrawal ∑WU to
annually available freshwater resources WAN from renewable groundwater as well as
from surface water and precipitation [14] (see Eq. 2.2):
P
W U W AG þ W I þ W D þ W RL
WTANR ¼ ¼ : ð2:2Þ
W AN W AN

WTANR ¼ withdrawal-to-availability ratio of natural renewable resources


∑WU ¼ annual withdrawal by users
WAN ¼ water available annually from natural renewable resources
WAG ¼ annual withdrawal by agriculture
WI ¼ annual withdrawal by industry
WD ¼ annual withdrawal for domestic use
WRL ¼ annual reservoir losses of withdrawal
2.2 Global Withdrawal and Consumption of Freshwater and Water Shortage 25

Table 2.4 Water stress indicators WTA and WSI and degree of water stress
Withdrawal-to availability (WTA) ratio/water stress Degree of water stress or
indicator (WSI) exploitation
0.3–0.5 Slight
0.5–0.7 Moderate
0.7–1.0 Heavy
>1 Severe, overexploited

By international agreement, the intensity of water stress or degree of natural water


resource exploitation is assigned to the following WTA/WSI values (Table 2.4).
When the withdrawal-to-availability (WTA) ratio is determined for different
regions of the world, it becomes apparent that large areas of the Earth’s surface are
exhibiting high levels of water stress (WTA/WSI higher than 0.7) (see Fig. 2.10).
More specifically, this concerns large areas of North Africa, the Middle East, the
Gulf Region, Southern Asia, northern and northwest China, the western USA, and
parts of the west coasts of South America and Australia. It is estimated that over
2 billion humans in more than 40 countries are currently affected by water shortage.
The major man-made factors leading to water shortages are of a demographic,
economic, social, and technological nature:

• Population growth and increasing per-capita water demand: as human per-capita


incomes rise, water use is increasingly seen as part of the quality of life. This is
resulting in higher specific water consumption for daily needs and recreational
activities in the affluent industrialized regions of the world.
• Urbanization: growing concentration of global population in centres of high
population density. It is estimated that while 48% of the world’s population in
2000 live in such population centres, this figure will rise to 60% by 2030. The
high water demand in such centres of population can no longer be met from
locally available natural resources or even from the more distant environs of such
megacities. In the majority of cases, the available groundwater reserves are being
exploited beyond their renewal potential and will, therefore, quickly become
exhausted.
• Economic growth and concomitant increase in industrialization: this can lead in
some cases to levels of water demand that can no longer be met by the natural
resources of the relevant region.
• Growing water demand of agriculture: population growth increasingly requires
the development of land for food production as well as the irrigation of that land
by “Blue Water”. Water use for artificial irrigation is consumptive, i.e. that water
is no longer regionally available for reuse.
• Pollution of surface water and groundwater by domestic and industrial effluents:
such freshwater resources either can no longer be used or can be reused only after
extensive treatment.
26
2

Fig. 2.10 Global view of existing freshwater shortage and water stress, 2009 (source: UNEP GRID ARENDAL, Philippe Rekacewicz, Smakthin, Revenga and
Döll, 2004)
Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .
2.2 Global Withdrawal and Consumption of Freshwater and Water Shortage 27

The causes of water shortage, therefore, are increasingly of an anthropogenic


nature. This is the case not only with regard to the above-listed factors but especially
also in connection with global warming, which involves a significant intervention by
humankind in the global water cycle.
Also from a global perspective, human withdrawal of exploitable water from the
Earth’s freshwater reserves is no longer on a negligible scale and is ever more
approaching the limits of the freshwater resources available to humankind. The
impacts of global warming on the global water cycle are likely to exacerbate this
situation rather than improve it.
Suitable models for predicting the future freshwater situation both globally and
regionally were developed or are in development [15, 16]. For determining the
global and regional availability of water, these models include factors that take
account of:

• Precipitation and evapotranspiration as well as the impact of global warming


• Groundwater resources and their renewal potential
• Surface water resources

With regard to the withdrawal and consumption of water, such models carry out
an analysis of demographic, economic, social, and technological trends, such as:

• Population growth and per-capita water demand


• Growth of urbanization
• Economic growth and industrialization
• Development of agricultural land and demand for artificial irrigation
• Losses of exploitable water resources through environmental influences
• Technological developments that extend the availability of freshwater resources
and improve the efficiency of water use (e.g. seawater desalination)

Figure 2.11 shows the results of such modelling for likely water stress in various
regions of the world in 2000 and 2050. The already existing water shortage situation
in 2000 in various regions of the world will, in some cases, worsen significantly by
2050. This applies particularly to parts of North Africa, South Africa, West Africa,
the Middle East, the Gulf Region, parts of southern Europe around the Mediterra-
nean Sea, the western USA, the east and west coasts of South America, southern
India, northern China, and Western Australia.
A sharp, sectoral rise in water scarcity and water stress is predicted, especially for
the world’s cities and megacities. The water supply shortages that already exist in
some cities and their urban catchment areas will worsen significantly in the coming
decades. Major reasons for this are contamination of surface water and groundwater
resources as well as overexploitation of aquifers, the latter additionally leading to
vegetation damage and soil subsidence owing to falling groundwater levels.
28 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.11 World water stress, at 2000 and predicted up to 2050 (source: Alamo, Flörke, Matovelle,
Menzel, Centre for Environmental Systems Research, University of Kassel, 2010)
2.3 Sustainable Water Management Systems and Their Components 29

2.3 Sustainable Water Management Systems and Their


Components

2.3.1 Supply Management

Water supply in most regions and countries of the Earth has to date usually followed
the principle of making the required quantities of freshwater available in line with
demand. Where industrial development, population growth, or agricultural expan-
sion gives rise to an increased water demand, the necessary additional supply
capacity is made available by:

• Increased usage of existing natural resources (increased withdrawal of surface


water and groundwater)
• Extension of existing/development of new natural resources, such as use of
higher-salinity groundwater resources or brackish water resources
• Upgrading of existing systems for collection and storage of precipitation (capac-
ity expansion of dams and reservoirs), i.e. conversion of “Green Water” into
accessible “Blue Water”.

The necessary infrastructure for such supply management is made up of:

• Systems for storage, buffering, and withdrawal of water, such as dams, reservoirs,
wells, and pumping stations as well as extraction and transmission piping
• Facilities for treatment of the withdrawn raw water according to its quality and
required standards of the treated product water
• Distribution systems for transport of water to consumers

The main purpose of such supply management is to ensure that the required
supply capacities can be reliably made available at all times by the existing infra-
structure in line with demand. With such a supply concept, the response to an
increase in water consumption is to improve the supply capacity through investment
in resources and to expand the capacity of withdrawal and treatment systems as well
as the distribution infrastructure. Frequently, however, it is accepted that natural
freshwater resources, especially groundwater, will be overexploited and that the
principles of sustainable water management will be violated. In countries in semi-
arid and arid zones with only few natural and easily exhaustible water resources,
supply management quickly reaches the point at which additional water demand can
no longer be met by an expansion of natural resources. In such cases, seawater
desalination is increasingly being used to supplement the water supply from existing
natural resources.
30 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

2.3.2 Water Demand Management [17, 18]

A broadening of supply management, and also an alternative to such a conventional


mode of management, is to extend the scope of water management to include
measures with an effect both on water consumption itself and also on the efficiency
of water consumption. This applies above all to those regions with only few or
difficult-to-access natural water resources or low or highly variable seasonal precipi-
tation frequency or to zones of high urbanization. Such a system of water demand
management also includes components aimed at sustainable management of existing
natural water resources. Both water management concepts can be combined to form
an integrated water management system capable of guaranteeing not only security of
supply but also more efficient water use and environmentally sustainable manage-
ment of natural resources [19, 20] (see Fig. 2.12).
The purpose of water demand management is to improve the efficiency of water
use and also to optimize and reduce water consumption. In addition, use of natural
water resources should focus more strongly on aspects of sustainability and environ-
mental compatibility. Furthermore, population growth and economic development
in a given region should not result in increased stress on natural water resources. The
degree of access to existing resources should be managed by appropriate control of
water consumption.

Water Demand Management


Water Supply Management
• Water consumption optimization
• Precipitation catchment (dams,
reservoirs) • Water usage efficiency improvement
• Water abstraction of surface water • Water losses minimization
and groundwater and transport (unaccounted-for water )
• Wastewater reuse and treated
• Water treatment wastewater distribution
• Seawater abstraction • Stormwater run-off catchment and
reuse
• Seawater desalination and
potabilization • Rainwater harvesting
• Product water transport and • Grey water reuse
distribution • Groundwater and aquifer recharge
• Product water quality management • Drought management
and control • Water cycles quality management
• Leakage detection and repair and control

Integrated
Water
Management

Fig. 2.12 Components and goals of integrated water management


2.3 Sustainable Water Management Systems and Their Components 31

Key measures for reducing consumption and improving water efficiency are:

• In the municipal water supply area


– Reduction of supply system losses (physical losses of unaccounted-for or
non-revenue water). Such water losses can account for up to 40% of the product
water fed into supply systems. In a well-maintained water supply system, such
losses make up between 10% and 20% of the water supply quantity.
– Addition of wastewater treatment facilities to municipal sewage treatment
plants and supply of treated wastewater to municipal and industrial consumers
with lower-quality requirements.
– Collection of stormwater for use as freshwater.
– Motivation of citizens to increase their collection and use of rainwater and grey
water.
– Reduction of per-capita water consumption through appropriate awareness-
raising campaigns about the value of water and its prudent use.
• In the industrial area
– Reduction of freshwater consumption in industrial processes through waste-
water reuse and wastewater recycling
– Initiation of product-oriented minimization of water consumption,
i.e. application of technologies with optimized consumption of process water
– Use of lower-quality water in appropriate production processes
• In the agricultural area
– Optimization of water consumption for irrigation of agricultural land through
application of efficient irrigation systems. Artificial irrigation in agriculture
withdraws the highest percentage of freshwater from the Earth’s natural
freshwater resources and is the dominant water consumer in most regions of
the world. Also, the majority of water use for artificial irrigation is consump-
tive, i.e. the water is regionally no longer available for reuse or multiple use.

Sustainable and environmentally compatible management of natural freshwater


resources comprises:

• Groundwater renewal in regions where the extent and frequency of precipitation is


not sufficient to prevent vegetation damage and soil subsidence through lowering
of the groundwater level because of unbalanced groundwater withdrawal
• Stringent standards on wastewater discharge quality in order to prevent pollution
of surface water that is used as a source of drinking water and process water
• Protection of groundwater against contamination by effluents and chemicals

2.3.3 Integrated and Sustainable Water Management [21]

The merging of the two water management concepts—one oriented towards con-
sumption and the other towards supply—to form a single integrated and sustainable
water management concept gives rise to a complex multidisciplinary operation and
organization system including technical, environmental, economic, social, and
demographic components.
32 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Precipitation
Seawater

Precipitation
catchment and Extraction
Surface water &
Groundwater fresh water
(Lakes, Rivers) storage (Dams, Transport
Reservoirs)

Extraction Desalination
&
Transport

Transport
Treatment and
Distribution
Transport
and
Distribution

Consumer

Agriculture Industry Municipality

Drinking
water
Wastewater
discharge Other
usages

Wastewater
Reuse Wastewater
Treatment discharge

Conventional water management


Sustainable water management

Fig. 2.13 Technical structure of an integrated water management system

Figure 2.13 presents the technical structure and configuration of such an


integrated water management system.
In a conventional water supply management system, freshwater is withdrawn
from existing natural resources, followed by appropriate treatment, and then sup-
plied to consumers in agriculture, industry, and municipalities. In the interests of
sustainable water management, an integrated water management system in addition
includes the treatment of wastewater from industrial and municipal consumers for
reuse. The treated wastewater can be used for irrigation in agriculture or in
municipalities for consumers with lower-quality requirements. Alternatively,
2.3 Sustainable Water Management Systems and Their Components 33

suitably high-grade treatment allows such water to be reused as drinking water.


Furthermore, industrial demand for process water can be met by such recycled water.
As an additional measure for sustainability, treated wastewater can be fed into
groundwater resources and aquifers for their protection and replenishment as well
as into dams and reservoirs in order to increase the water reserves.
However, the water recycling components of an integrated water management
system consist not only of suitable treatment facilities. Frequently, additional collec-
tion, transport, and distribution systems are required for supplying the treated
wastewater to consumers separately from the normal drinking water supply. Conse-
quently, the technical structure of an integrated water management system is signifi-
cantly more complex than in a conventional water supply system. The same is true of
capacity management and quality control. Water supply and wastewater disposal
must be closely interlinked both organizationally and also in relation to information
interchange so as to allow efficient processes within an integrated management
system. Technical measures must be supported by appropriate management and
organization systems that not only provide a comprehensive overview of the actual
operation situation within the management network but also take account of external
influences and allow predictions with regard to appropriate operational and structural
adjustments.

2.3.3.1 Wastewater Reuse


Wastewater reuse can be based on both municipal and industrial wastewater. How-
ever, the water used by the biggest consumer of freshwater, which is agriculture, is to
all intents and purposes unavailable for reuse. The majority of water use for artificial
irrigation of agricultural land is consumptive, i.e. it is lost to the local or regional
water balance. Thus, although agriculture is a user of both freshwater and also
recycled wastewater, it makes hardly any or no contribution whatsoever to water
reuse. Consequently, the ratio of agricultural water use to a region’s total freshwater
consumption will have a key influence on the extent to which that region’s freshwa-
ter consumption can be reduced by water reuse. The total amount of wastewater
available for reuse (∑WWRUA), therefore, is made up of the accessible wastewater
from industries WWRI and municipalities WWRD (Eq. 2.3):
X
WW RUA ¼ WW RI þ WW RD : ð2:3Þ

∑WWRUA ¼ wastewater accessible for reuse


WWRI ¼ industrial wastewater accessible for reuse
WWRD ¼ domestic wastewater accessible for reuse

The available amount of wastewater WWRI, D available for treatment in reuse


facilities in each case is the individual consumer’s freshwater withdrawal less its
consumptive water use less the water losses in the supply network between point of
withdrawal and point of consumption (Eq. 2.4):
34 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

 
WW RI,D ¼ W I,D  1  f CI,D  f LUFWI,D  f AI,D : ð2:4Þ

WWRI, D ¼ industrial or domestic wastewater accessible for annual reuse


WI, D ¼ annual withdrawal by industry or domestic users
fCI, D ¼ consumption ratio of industrial or domestic users
fLUFWI, D ¼ water loss ratio (rate of physical losses of unaccounted-for water (UFW))
for industrial or domestic users
fAI, D ¼ wastewater accessibility factor for industrial or domestic users

Consumptive water use refers to the amount of water that stays with the industrial
or municipal consumer and is not discharged as wastewater. Industrial consumptive
usage is highly dependent on the relevant branch of industry, type of water supply,
production process, and climatic conditions. In most industries, consumptive water
usage is between 5% and 20%, although it can also be up to 30% or 40%. Municipal
consumptive water usage is made up mainly of leakage losses in supply networks
(physical water losses of unaccounted-for water (UFW)), evaporation, and watering
of lawns and parks in recreational areas. Municipal consumptive water usage can be
between 5% and 60% of freshwater withdrawal and is highly dependent on the
quality and maintenance of water supply networks as well as on climatic conditions.
Additional factors that determine the possible extent of wastewater reuse are the
water efficiency of consumers as well as the degree to which their wastewater is
collected. The only wastewater that can be treated for reuse are those discharged by
municipal or industrial users for collection and do not otherwise escape a controlled
collection.
The amount of treated wastewater ultimately available for reuse depends also on
the efficiency of the wastewater reuse treatment processes, i.e. on the efficiency of
conversion of untreated wastewater to product water (Eq. 2.5):

W RUI,D ¼ WW RI,D  f REI,D : ð2:5Þ

WRUI, D ¼ treated wastewater for reuse from industry or domestic users


fREI, D ¼ water reuse efficiency for industrial or domestic users

The amount ∑WRU of reusable water from the treatment of available municipal
and industrial wastewater is calculated from the total treated wastewater of both
groups of consumers with, respectively, the applicable factors and efficiencies for:

• Consumptive water usage


• Per-capita water consumption
• Water losses in the municipal water distribution network and, for industrial
consumers, in their internal supply systems
2.3 Sustainable Water Management Systems and Their Components 35

• Degree of wastewater collection from municipal and industrial consumers


• Efficiency of wastewater treatment processes

See Eqs. (2.6) and (2.7):


X
W RU ¼ W RUI þ W RUD : ð2:6Þ
X
W RU ¼ W I  ð1  f CI  f LUFWI Þ  f AI  f REI þ W D
 ð1  f CD  f LUFWD Þ  f AD  f RED : ð2:7Þ

∑WRU ¼ total reused water from wastewater of industry and domestic users

These algorithms combined with the choice of common values for the above-
mentioned factors allow an estimation of the possible extent of wastewater reuse for
regions according to their structures of settlement and industrialization (predomi-
nantly rural, mixed rural/urban, predominantly urban). The benchmark for structural
rating is the percentage of total freshwater withdrawal accounted for by agriculture,
industry, and municipalities in each individual region (Table 2.5).
The results of these calculations show that the highest percentage of wastewater
reuse is possible in a predominantly urban environment. The reason for this is that,
on account of the high population density and—in comparison with rural regions—
shorter distances, the organizational and technical conditions necessary for waste-
water reuse are easier to implement. More specifically, these necessary conditions
include the minimization of freshwater losses in the water supply network and the
establishment of a separate system for supplying the treated wastewater to
consumers. A water reuse rate of between 40% and 50% appears possible in urban
areas, whereas, in exclusively rural regions, a maximum of 10% seems likely, while,
in mixed rural/urban areas, a maximum of 20% wastewater reuse might be achiev-
able. This depends on the degree to which the above-mentioned factors can be
optimized in each individual case. A reduced per-capita water consumption, how-
ever, will result in a decreased amount of municipal wastewater available for
wastewater reuse. The same applies to optimization of specific freshwater consump-
tion in industry and the quantity of industrial wastewater available for wastewater
reuse.
Wastewater reuse alone, therefore, cannot make up for an existing freshwater
deficit in a region or municipality. An additional freshwater supply from external
natural or artificial resources is necessary.
36 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Table 2.5 Regional structure and resulting potential for wastewater reuse (data source:
UNESCO—International Hydrological Program—World water resources and their joint use;
I.A. Shiklomanov, State Hydrological Institute (SHI), St. Petersburg, 1999)
Rural/
Global Rural urban Urban
Parameter Symbol Dimension average dominating mixed dominating
Withdrawal rate, WAG % of ∑WU 66 80 60 5
agriculture
Wastewater reuse, industry
Withdrawal rate, WI % of ∑WU 9 5 20 50
industry
Consumption ratio, fCI – 0.15 0.20 0.30 0.40
industry
Water loss part of fLUFWI – 0.15 0.20 0.15 0.10
UFW ratio,
industry
Wastewater fAI – 0.80 0.70 0.85 0.90
accessibility factor,
industry
Water reuse fREI – 0.80 0.80 0.80 0.85
efficiency, industry
Water from WRUI % of ∑WU 4.0 1.7 7.5 19.1
industry reuse
Wastewater reuse, domestic
Withdrawal rate, WD % of ∑WU 20 10 17 40
domestic
Consumption ratio, fCD – 0.11 0.20 0.15 0.10
domestic
Water loss part of fLUFWD – 0.25 0.30 0.20 0.10
UFW ratio,
domestic
Wastewater fAD – 0.70 0.60 0.80 0.90
accessibility factor,
domestic
Water reuse fRED – 0.70 0.70 0.75 0.80
efficiency,
domestic
Water from WRUD % of ∑WU 6.3 2.1 6.6 23.0
domestic, reuse
Water from reuse ∑WRU % of ∑WU 10.3 3.8 14.1 42.2
rate of industry
and domestic total

2.3.3.2 Seawater Desalination


Where the existing capacity of natural freshwater resources is insufficient, seawater
desalination can be used to provide additional capacity, either as an alternative or as
an addition to conventional water supply management measures.
2.3 Sustainable Water Management Systems and Their Components 37

Depending on local or regional conditions, seawater desalination can be used to


meet water demand that cannot be covered by existing natural resources or by
additional consumption management measures, such as wastewater reuse, rainwater
collection, and increased stormwater usage. In such cases, the necessary additional
capacity is made available by seawater desalination.
Alternatively, however, as is already the case in many highly populated semi-arid
regions of the world, seawater desalination can serve as a region’s principal source of
freshwater. This is particularly the case where a region’s natural resources, espe-
cially its groundwater reserves, are already to a large extent exhausted by overex-
ploitation or are nearing exhaustion.
An alternative option for seawater desalination, particularly from the perspective
of sustainable water management, is to feed the product water from such plants into
existing groundwater resources and aquifers. This can be done in the same measure
as water is withdrawn by supply management so as to achieve a balance between
withdrawal and replenishment. Alternatively, such feed-in can serve to reverse
damage to groundwater resources. The use of aquifers for intermediate storage of
water, to buffer seasonal variations in natural water resources, can also be part of a
sustainable water management regime. Another positive aspect of such a policy is
that a desalination plant need not be adjusted in terms of capacity management to suit
seasonal variations in precipitation or in available amounts of natural water but can
be kept in continuous operation. The same positive aspect applies where surplus
product water from seawater desalination is fed into existing dams or reservoirs.

2.3.3.3 Freshwater Resources in Integrated Water Management


and Their Allocation
In an integrated water management system, freshwater resources are made up of
(Fig. 2.13):

• Withdrawal from natural water resources (surface water and groundwater, as


sustainably as possible)
• Wastewater reuse
• Seawater desalination

Natural water resources are divided into those that are renewable by precipitation
and those that are non-renewable, so-called “fossil” resources (Eq. 2.8):
X
W N ¼ W ANR þ W NNR : ð2:8Þ

∑WN ¼ water annually available from natural resources


WANR ¼ annual freshwater available from renewable natural resources
WNNR ¼ water withdrawal from natural non-renewable resources
38 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

The degree to which a country’s or region’s natural renewable freshwater


resources are strained by withdrawal is characterized by the withdrawal-to-availabil-
ity ratio WTANR (see Eq. 2.2). Where WTANR is greater than 1, the water supply
uses freshwater from non-renewable sources in addition to withdrawal from renew-
able water resources. Where water is withdrawn exclusively from natural resources,
the extent of available renewable freshwater is calculated using the WTANR ratio
according to Eq. (2.9):
P
W UN
W ANR ¼ : ð2:9Þ
WTANR

WTANR ¼ withdrawal-to-availability ratio of natural renewable resources


∑WUN ¼ annual freshwater withdrawal by users from natural resources

Withdrawal from non-renewable resources is then calculated according to


Eqs. (2.10) and (2.11):

W NNR ¼ W ANR  ðWTANR  1Þ: ð2:10Þ


X  
1
W NNR ¼ W UN  1  : ð2:11Þ
WTANR

In an integrated water management system, the withdrawal of freshwater from


natural resources WNR is supplemented by provision of water from artificial,
man-made sources. These sources are water reuse or wastewater recycling and
seawater desalination (Eq. 2.12).
X X X
W U = W NR þ W A ¼ W NR þ W SD þ W RU : ð2:12Þ

∑WU ¼ total water withdrawal rate by users


WNR ¼ water from natural resources
∑WA ¼ freshwater from artificial sources
WSD ¼ water from seawater desalination
∑WRU ¼ water from wastewater reuse

For sustainable water management, however, natural freshwater should be with-


drawn exclusively from renewable resources, and the amount withdrawn should not
place undue stress on those resources (see Table 2.3). To mitigate water stress, some
of the freshwater withdrawn from natural resources can be replaced by water from
seawater desalination and wastewater reuse. The amount WNRS that can still be
withdrawn from natural renewable resources WARN can be defined by the ratio
WTANS and can be calculated using this parameter (Eqs. 2.13 and 2.14):
2.3 Sustainable Water Management Systems and Their Components 39

W NRS
WTANS ¼ ð2:13Þ
W ANR
W NRS ¼ W ANR  WTANS : ð2:14Þ

WTANS ¼ WTA from natural sustainable resources


WNRS ¼ sustainable water withdrawal from natural resources

Similar to a supply management system based exclusively on withdrawal from


natural freshwater resources, a withdrawal-to-availability ratio WTAS that defines
the ratio of freshwater withdrawal to available resources, including artificial
resources, can also be defined for an integrated water management system
(Eq. 2.15). With a withdrawal-to-availability ratio WTAS ¼ 1, only withdrawal for
consumption is covered. If WTAS < 1, additional capacity from artificial sources is
also available, such as for groundwater renewal:
P P
WU WU
WTAS ¼ ¼ P
W TS W NRS þ W SDS þ W RU
P
WU
¼ P : ð2:15Þ
WTANS  W ANR þ W SDS þ W RU

WTAS ¼ WTA sustainable


WTS ¼ water withdrawal rate sustainable resources

If, in such a sustainable integrated system of water management, some of the


water withdrawal from natural resources is substituted by withdrawal from artificial
WTANS
sources, the quotient WTA NR
serves as a measure of the residual water stress on natural
resources (Eq. 2.16):
P
WTANS 1 W þ W RU
=  SDSP : ð2:16Þ
WTANR WTAS WU

The amount WSDS of water that must be made available by seawater desalination
for a given water consumption ∑WU and a given degree of wastewater reuse ∑WRU
WTANS
in order to achieve the required usage rate of renewable freshwater resources WTA NR
or a correspondingly reduced level of water stress is calculated from Eq. (2.17):

X   X
1 WTANS
W SDS ¼ WU    W RU : ð2:17Þ
WTAS WTANR

WSDS ¼ water from seawater desalination for sustainable freshwater management


40 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

1.0
WTANS/WTANR
Usage rate of 0.9
renewable water WTAS = 0,8
available Reuse rate 5%
0.8 Reuse rate 10%
Reuse rate 20%
0.7 Reuse rate 30%
Reuse rate 40%

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0% 20% 40% 60% 80% 100% 120%
Seawater desalination capacity as a percentage of total water withdrawal

Fig. 2.14 Required seawater desalination capacity versus usage rate of natural freshwater
resources (data source: FAO Aquastat; data of years 2000–2006)

Depending on the required value for WTAS, provision is made for additional
production capacity from artificial resources for feeding into aquifers for the purpose
of replenishment and/or intermediate storage.
Figure 2.14 presents the required amount of product water from seawater desali-
WTANS
nation for different usage rates WTA NR
of natural freshwater resources, wastewater
reuse rates of 5–40%, and WTAS of 0.8.
To reduce the usage rate of natural renewable water resources from 100%
(equivalent to high water stress) to 50% (equivalent to moderate stress), it is
necessary, depending on the water reuse rate, for seawater desalination to make up
between 35% and 70% of total water consumption.
In dry, arid and semi-arid zones of the world, however, insufficient renewable
freshwater resources are available to supply the population, agriculture, and industry
exclusively from that source. This is especially true in the high-growth regions of the
Persian Gulf and North Africa, where freshwater withdrawal from natural resources
is based mainly on the use of non-renewable groundwater resources.
Table 2.6 shows the following for some MENA countries:
Table 2.6 Water stress on natural resources for different countries (data source: FAO Aquastat; data of years 2000–2006)
Water
resources Fresh Groundwater Water Fresh water over-
renewable water resources Groundwater withdrawal Groundwater Withdrawal- to- exploitation rate
actual withdrawal renewable withdrawal total exploitation availability ratio OER
WGRW * 100/ WTANR = ∑WN/ OER = (∑ WN
WAN ∑ WN WGR WGRW WT WGR WAN WAN) * 100/WT
Country 109 m3/y 109 m3/y 109 m3/y 109 m3/y 109 m3/y % – %
Kuwait 0.02 0.49 0.02 0.42 0.91 2075 24.7 51.8
United 0.15 3.05 0.12 2.10 4.00 1750 20.3 72.5
Arab
Emirates
Saudi 2.40 22.64 2.20 21.37 23.67 971 9.4 85.5
Arabia
Libya 0.60 4.33 0.50 4.31 4.33 862 7.2 86.2
Qatar 0.06 0.26 0.06 0.22 0.44 381 4.4 45.9
Bahrain 0.12 0.26 0.11 0.24 0.36 213 2.1 37.8
Yemen 2.10 3.54 1.50 2.40 3.54 160 1.7 40.7
Israel 1.78 1.81 1.23 0.95 2.80 78 1.0 1.2
2.3 Sustainable Water Management Systems and Their Components

Jordan 0.94 0.93 0.54 0.55 1.51 102 1.0 0.0


Oman 1.40 1.21 1.30 1.21 1.32 93 0.9 0.0
Tunisia 4.60 2.84 1.60 1.90 2.85 119 0.6 0.0
Algeria 11.67 6.14 1.52 0.77 6.16 50 0.5 0.0
Spain 111.50 32.40 29.90 5.70 32.36 19 0.3 0.0
41
42 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

• Total water withdrawal


• Available renewable natural freshwater resources
• Freshwater withdrawal from natural resources
• Extent of renewable groundwater resources

Based on calculations using these parameters, this table presents the WTA ratio,
groundwater usage rate, and freshwater overexploitation rate (OER) as criteria for
evaluating the water stress in each country. The OER value expresses freshwater
withdrawal from non-renewable natural resources as a percentage of total water
withdrawal for each country (Eq. 2.18):
P
W W þW þ W RU
OER ¼ PNNR ¼ 1 2 NRS PSD : ð2:18Þ
WU WU

OER ¼ overexploitation rate

The table also takes into consideration the extent to which non-natural sources,
i.e. water reuse and seawater desalination, contribute to a country’s water supply.
According to these statistical water data, which originate from 2000 to 2008, all
Gulf countries and Libya exhibit severe to extreme water stress in terms of natural
water resources. This is due in all cases to significant overexploitation of
non-renewable groundwater resources. The overexploitation rate (OER) provides a
more realistic impression of overexploitation of natural water resources than the
WTA value or water stress indicator (Fig. 2.15).

Kuwait

United Arab Emirates

Saudi Arabia
Withdrawal-to- Availability Ratio

Libya
Overexploitation rate [Extraction
from nonrenewable sources as % of
total withdrawall]
Qatar

Bahrain

Yemen

Israel

0 10 20 30 40 50 60 70 80 90

Fig. 2.15 Water stress and overexploitation of non-renewable freshwater resources for different
countries (data source: FAO Aquastat; data of years 2000–2006)
2.3 Sustainable Water Management Systems and Their Components 43

As indicated by the OER value, freshwater from non-renewable natural resources


as a proportion of total water consumption in these regions is in the range 40–80%.
The OER value, however, does not follow the value of the water stress indicator.
Some countries with a significantly lower WTA than other countries are shown
according to their OER to have noticeably higher overexploitation of their natural
water resources than is suggested by the WTA value.
The high rate of overexploitation of natural resources in the high-growth dry
regions of the Persian Gulf and North Africa will in the long term make it necessary
for these regions’ dependency on non-renewable freshwater resources to be reduced
even more significantly than before through the use of artificial freshwater sources.
Without such measures, it will not be possible to guarantee the future supply of water
in these countries, assuming that population figures continue to rise and that there is
further growth of industry and agriculture.
Using Eq. (2.19) and with knowledge of the available natural freshwater
resources ∑WN and the quotient of their usage WTA
WTANS
NR
, it is possible to create a profile
for a given region or country that indicates the extent to which artificial freshwater
resources ∑WA from seawater desalination and water reuse need to be made avail-
able in order to suitably reduce overexploitation of the existing renewable water
resources (characterized by the OER ratio):
X X WTANS X
W SD ¼ W U  ð1  OERÞ  WN  2 W RU: ð2:19Þ
WTANR
Figure 2.16 presents such a profile for a region in which only 10% of total water
withdrawal can be met from renewable resources.

95%

Percentage of WW reuse rate 0 %


85%
Desalination capacity
[% of Withdrawal] WW reuse rate 10 %

75% WW reuse rate 20 %


WW reuse rate 30 %

65% WW reuse rate 40 %

55%

45%

35%

25%

15%

5%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90%
Overexploitation rate [%]

Fig. 2.16 Effect of seawater desalination rate on overexploitation rate of natural water resources
(data source: FAO Aquastat; data of years 2000–2006)
44 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Kuwait

Qatar

Bahrain
Artifical Freshwater sources rate
[% of Total Withdrawal 2000 -
United Arab Emirates
2006]
Desalination capacity [% of Total
Oman
Withdrawal rate 2000 - 2006]

Saudi Arabia Waste water reuse capacity [ %


of Total Withdrawal 2000 - 2006]
Israel

Algeria

Libya

Spain

Jordan

Tunisia

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 [%]

Fig. 2.17 Artificial freshwater withdrawal, desalination, and reuse capacity for different countries

Especially in the Gulf states, the shortage of natural water resources has led to
significant growth of artificial freshwater production capacities (Fig. 2.17).
Compared with the above-presented statistical water data, more recent surveys of
the existing capacities for seawater desalination and water reuse even demonstrate
that the installed plant capacities for artificial freshwater production significantly
exceed the previous value for total water withdrawal in some of these countries.
Although this may be due to a substantial increase in water consumption in the
meantime, which has led to this increased capacity expansion, it is also possible that
this additional plant capacity is being used to suitably reduce the withdrawal of
freshwater from non-renewable sources and thus make water management in these
countries more sustainable.
Capacity expansion in terms of artificial freshwater production in these regions is
clearly focused on seawater desalination. The reasons for the less widespread
application of water reuse and wastewater recycling, particularly in the municipal
sector, may, as already mentioned above, include the more complex technical and
organizational structure of such systems in rural and extensive regions. Another
reason is surely the fact that water reuse is restricted mainly to agricultural irrigation
and that treated wastewater is used predominantly for applications with lower-
quality requests.
In dry, semi-arid and arid zones of the world, therefore, seawater desalination
already today plays a vital role in the supply of water to municipalities, industries,
and agricultural sectors. The significance of seawater desalination will continue to
grow in future in cases where the aim is to stop overexploitation of natural resources
leading to permanent environmental damage or to ensure that a deteriorating water
2.4 Seawater Desalination Processes and Their Development 45

quality does not result in such water resources becoming increasingly difficult or
even impossible to use. Seawater desalination and water reuse or wastewater
recycling are essential elements of an integrated and sustainable water management
system. Water reuse alone is no substitute for seawater desalination and can only
supplement it. A minimum freshwater infeed into water supply systems and water
cycles is necessary in order to maintain water quality and to assure security of supply
in line with demand.
However, the sustainability of seawater desalination is determined also by the
manner in which the required process energy is generated. It is currently still the case
that most of the necessary heat and electricity come from power plants fired by
fossil-based, non-renewable energy sources. However when seawater desalination is
coupled with power generation from renewable energy sources, such as wind,
photovoltaic, and solar thermal, this considerably improves the sustainability and
ecology of integrated water supply systems with seawater desalination also in terms
of their energy supply.

2.4 Seawater Desalination Processes and Their Development

2.4.1 Types of Desalination Processes and Their Configurations

The seawater desalination process breaks down saline seawater into a low-salinity
product stream and a higher-salinity concentrate. Of the available desalination
processes, the two that are today used almost exclusively in commercial large-
scale applications are the technologies of thermal desalination (evaporation/distilla-
tion) and reverse osmosis membrane separation.
In the thermal distillation process, seawater is evaporated and the resulting vapour
is condensed to produce low-salinity product water. Phase separation requires the
process to be supplied with an appropriate amount of heat. Electrical energy is also
needed to move the mass streams (seawater, brine, and distillate) through the
desalination system and, to some extent, also to generate the vacuum necessary for
operation.
In the reverse osmosis process, desalination is by means of a semipermeable
membrane that rejects most of the salt contained in the seawater. The driving force
behind the desalination process is the pressure at which the reverse osmosis mem-
brane is supplied with seawater. Electrical energy is required in order to generate the
pump pressure necessary for this purpose.

2.4.1.1 Thermal Processes


Thermal processes can be subdivided into two inherently different evaporation
processes, namely:

• Multi-stage flash evaporation (MSF)


• Multi-effect distillation (MED)
46 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Depending on its technical configuration, the MED process can, in turn, be


further subdivided into:

• MED without vapour compression (plain MED)


• MED with thermal vapour compression (MED-TVC)
• MED with mechanical vapour compression (MED-MVC)

In multi-stage flash evaporation (MSF) (see Fig. 2.18), heated seawater is


passed through an evaporator unit with serially arranged evaporation chambers.
Superheating of the water in combination with decreasing pressure in each of the
individual chambers results in spontaneous steam formation (flashing). Whereas
there is a fall in temperature in the chambers between the first stage (which has the
highest operating temperature) and each of the following evaporation stages, the salt
concentration in the water rises due to evaporation. The resulting vapour is
condensed in heat exchangers located in the upper part of each flash chamber. The
vapour transfers its heat to the colder seawater that is fed into the MSF process and
which flows through these heat exchangers and is thereby heated. Before the vapour
condenses on the tube bundles of the heat exchangers, it is passed through demisters,
which reduce the salt concentration in the vapour, with the consequence that the
resulting distillate has a very low salt content. The distillate is collected in distillate
trays and removed in the final lowest temperature stage of the evaporator. Having
already been heated in the individual evaporation chambers, the seawater is provided
with the rest of the process heat needed for operation of the evaporator in the brine
heater, which is supplied with process steam. The brine that forms as the salt
concentration increases from stage to stage of the evaporator is also removed in
the final stage of the evaporator. In the most common configuration of the MSF
process, the brine is repeatedly circulated through the evaporator by a special brine
recirculation pump.
Multi-effect distillation (MED) (see Fig. 2.19) also is a multi-stage evaporation
process. In this case, however, evaporation is on the surfaces of heat exchanger tube
bundles and not by flashing. Each of the evaporation stages is equipped with such
heat exchangers, onto which the seawater is sprayed to allow it to evaporate.
Evaporation takes place from a liquid film that forms on the surfaces of the tube
bundles of the heat exchanger. Having passed through a demister, the vapour
produced in each stage is supplied as heating steam to the next stage of the MED
process, where the steam enters the interiors of the tubes of the heat exchanger. In
doing so, the steam cools down while at the same time transferring its heat to the
seawater, which evaporates on the outside of the heat exchanger bundle. The thus
formed condensate in the heat exchanger bundles of the individual stages is collected
and removed from the evaporator as distillate. Also with this process, the tempera-
ture and pressure in the individual stages (effects) decrease from one stage to the
next, whereas the salt concentration of the resulting brine increases. The residual
process heat is used in a final condenser to preheat the colder seawater feed. The
necessary process heat is supplied to the evaporator in the first, hot evaporator stage
by low-pressure process steam.
2.4 Seawater Desalination Processes and Their Development 47

Fig. 2.18 Diagram of MSF process


48 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.19 Diagram of plain MED process


2.4 Seawater Desalination Processes and Their Development 49

MED-TVC (see Fig. 2.20) is basically a MED evaporation process of the kind
described above. In this case, however, a thermo-compressor is used to supply
process heat and to compress the vapour. Vapour is withdrawn from the
low-temperature MED stages, compressed in the thermo-compressor by the injection
of medium-pressure steam and fed back into the first, hot MED stage. This process
makes it possible to reduce the number of MED stages (effects) while at the same
time improving the thermal efficiency of the process.
In the MED-MVC process, vapour compression is by means of a mechanically
driven compressor instead of a thermo-compressor. The evaporator system is other-
wise similar in principle to the above-described MED and MED-TVC processes. A
MED-MVC plant uses electrical energy for operation, with process heat being
needed only for starting the evaporator. Also, unlike the other thermal processes,
MED-MVC normally requires no additional cooling water.

2.4.1.2 Membrane Processes


The core process of a membrane desalination plant (see Fig. 2.21) is the reverse
osmosis membrane stage. The reverse osmosis process employs a semipermeable
membrane, which separates the saline seawater into a low-salinity product and a
higher-salinity concentrate. The driving force behind this separation process is the
pressure applied to the reverse osmosis membranes. The residual pressure of the
concentrate is utilized by energy recovery systems to increase the pressure of
seawater feed into the reverse osmosis stage.
To protect the surfaces of the membranes against deposits of seawater impurities,
which would adversely affect the salt-retaining properties and permeability of the
membranes, it is necessary for reverse osmosis desalination processes to be equipped
with appropriate pretreatment stages. Unlike in the case of a thermal desalination
plant, in a desalination plant of this type the required quality of the product water
also has an influence on the construction and configuration of the reverse osmosis
part of the desalination process.
Table 2.7 compares the MSF and MED thermal processes as well as seawater
reverse osmosis (SWRO) with regard to kind of energy required for operation,
energy consumption, and other process-specific design and operation parameters.
The MSF and MED thermal processes require not only process heat but also
electrical power as energy for their operation. The two processes differ in terms of
operating temperature, which, at a maximum of 114  C, is higher for the MSF
process than for the MED process, which has a maximum operating temperature of
70  C. MSF and MED-TVC are similar with regard to steam conditions. Plain MED
can be operated with process steam at a significantly lower vapour pressure or even
with hot water.
The performance ratio (PR) is a key indicator of an evaporator’s efficiency of heat
utilization. The higher the numerical value, the more efficient is the evaporator’s
utilization of heat for each quantity of distillate produced. The specific heat con-
sumption SECh of a thermal desalination process is calculated from the performance
ratio (PR) according to Eq. (2.20):
50 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Fig. 2.20 Diagram of MED-TVC process


2.4 Seawater Desalination Processes and Their Development

Fig. 2.21 Diagram of reverse osmosis seawater desalination process


51
52 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Table 2.7 Technical data and design parameters of MSF, MED, and RO desalination processes
Desalination processes
Membrane
Thermal processes processes
Multi-stage Multi-effect distillation MED
flash
distillation MED MED- MED-
Parameter Unit MSF plain TVC MVC SWRO

Operating C 90–114 55–70 55–70 55–70 <45
temperature
Energy supply
Kind of energy – Heat and electricity Electricity
Heat conditions bar Process steam Process Process
(without ejector 2.5–3.5 steam steam
steam) 0.25– 2.5–3.5
0.35
Hot
water
Specific energy consumption of desalination process (SWRO with pretreatment)
• Performance kg 8–10 8–10 8–15 – –
ratio PR product/
2326 kJ
• Heat SECh MJ/t 291–233 291– 291– – –
product 233 155
• Electrical kWh/t 11–16 4–6 8–16 – –
energy product
equivalenta
SECeq
• Electricity kWh/t 3.0–4.0 1.5–2.0 1.5–2.0 7.5– 3.0–5.0
SECe product 14.0
• Total SECt kWh/t 14.0–20.0 5.5–8.0 9.5– 7.5– 3.0–5.0
product 18.0 14.0
Other plant parameters
Recovery rate Y % ~30–50
Max. (without % ~30 ~50 –
cooling water)
YEvapor
Range (with % ~10–18 – –
cooling water)
Yprocess
Concentration – ~1.4–2.0
factor CF
Max. (without – ~1.4 ~2.0 –
cooling water)
CFEvapor
Range (with – ~1.1–1.2 – –
cooling water)
CFprocess
(continued)
2.4 Seawater Desalination Processes and Their Development 53

Table 2.7 (continued)


Desalination processes
Membrane
Thermal processes processes
Multi-stage Multi-effect distillation MED
flash
distillation MED MED- MED-
Parameter Unit MSF plain TVC MVC SWRO
Feed TDS range mg/l 30,000–100,000 1000–
50,000
Product TDS mg/l <10–30 <100a–<
range 500b
Actual max. plant and unit size
Unit/train size m3/day 80,000 13,300 68,000 3000– 25,000
5000
Plant size m3/day 880,000 40,000 800,000 17,000 500,000
a
Two-pass RO
b
Single pass RO

2326
SECh ¼ : ð2:20Þ
PR

SECh ¼ specific heat consumption [kJ/kg of product]


PR ¼ performance ratio

To be able to compare the energy consumption of thermal processes and mem-


brane processes, the electrical energy equivalent SECeq is determined for the heat
portion of total energy consumption of a thermal desalination process. This indicates
how much electrical power could be generated by a condensing power plant from the
heat required for desalination. The value of this parameter is dependent on the
necessary operating conditions (pressure and temperature) of the process steam of
the evaporator as well as on the conditions assumed for the condensing turbine of the
power plant. The values in Table 2.7 are based on a condensation temperature of
45  C, corresponding to a pressure in the condenser of 0.1 bar.
The total specific energy consumption SECt of a thermal desalination process is
calculated from the specific electrical energy consumption SECe and the electrical
energy equivalent SECeq according to Eq. (2.21):

SECt ¼ SECeq þ SECe : ð2:21Þ

SECt ¼ total specific energy consumption [kWh/tp]


SECeq ¼ specific equivalent energy consumption [kWh/tp]
SECe ¼ specific electrical energy consumption [kWh/tp]
54 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

With regard to the product recovery (yield) Y of a thermal process, a distinction


must be made between the yield Yevap of the evaporator itself and that of the process
Yproc (Eqs. 2.22 and 2.23):

F p  100
Y proc ¼ : ð2:22Þ
F ft
F p  100
Y evap ¼ : ð2:23Þ
Fm

Yproc ¼ yield of process [%]


Yevap ¼ yield of evaporator [%]
Fp ¼ flow product/distillate
Fft ¼ total feed flow
Fm ¼ makeup water (evaporator) feed flow

The seawater feed flow Fft to the thermal desalination process includes not only
the water fed into the evaporation process Fm but also the cooling water Fcw required
for removing the residual heat from the heat reject section of an MSF process or from
the final condenser of an MED process (see Eq. 2.24):

F ft ¼ F cw þ F m ð2:24Þ

Fcw ¼ cooling water feed flow


Fm ¼ makeup water (evaporator) feed flow

The discharge flow Fd of a thermal desalination process is then calculated


according to Eq. (2.25):

F d ¼ F ft  F m þ F cd : ð2:25Þ

Fcd ¼ concentrate/brine discharge flow


Fd ¼ process discharge flow

The corresponding concentration factors CFproc and CFevap are then also calcu-
lated from the two different product yields Yproc and Yevap (see Eqs. 2.26–2.29):

F ft
CFProc ¼ : ð2:26Þ
Fd
Fm
CFevap ¼ : ð2:27Þ
F cd
2.4 Seawater Desalination Processes and Their Development 55

1
CFproc ¼ Y proc : ð2:28Þ
1  100

1
CFevap ¼ Y evap : ð2:29Þ
1  100

CFevap ¼ concentration factor of evaporator


CFproc ¼ concentration factor of process

These concentration factors can be used to calculate the salt concentration of the
brine cb discharged from the evaporator as well as the salinity cd of the overall
discharge of the thermal desalination process (Eqs. 2.30 and 2.31):

cb ¼ cf  CFevap : ð2:30Þ

cd ¼ cf  CFproc : ð2:31Þ

cf ¼ salt concentration of feed into process


cd ¼ salt concentration of process discharge
cb ¼ salt concentration of brine discharge of evaporator

Owing to the low salt content of the distillate of a thermal process (see Table 2.7),
the amount of salt leaving the distiller within this part stream can be neglected.
According to their kind of energy supply and specific operating conditions, the
different seawater desalination processes can be variously configured with energy
supply systems (Fig. 2.22).
Thermal processes, which require both heat and electrical power for operation,
can be configured with energy supply systems that:

• Cover only the heat and power demand of the desalination process (self-
supplying systems with boiler and steam turbine for the internal power demand
of the plant)
• Cover only the heat demand of the desalination process, with the required
electrical power being imported (only boiler for process steam)
• Not only cover the heat and power demand of the desalination process but also
generate additional power for export (seawater desalination and power plants,
dual-purpose plants)

Where thermal seawater desalination is coupled with industrial facilities that


generate steam and power for their own production processes, the energy needs of
the desalination process are met also from the process energy supply (system with
imported heat and power). This allows low-pressure steam (which is unsuitable as a
56 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

Membrane
Thermal Processes
Processes

Desalination Reverse
Multistage Multieffect
Process Osmosis
Flash Distillation
RO
MSF MED

MED MED
MED
TVC MVC

Kind of
Energy Heat
Supply
Electricity

Type of
Single Purpose Dual Purpose Dual Purpose Dual Purpose Single Purpose
Desalination
Stand-alone Thermal Power and Water Power and Water Power and Water Stand-alone
Configuration Thermal Plant Hybrid Plant Membrane Plant Membrane Plant
Plant

Type of Energy
Self Supplying Self Supplying System with Heat System with Heat
Supply System with Power
System System with Supply and Power and Power
Configuration Import
Power Export Import Import

Fig. 2.22 Possible desalination and power generation configurations

source of process heat) to be advantageously utilized (especially for the MED


process).
For seawater desalination processes that require only electrical power for their
operation (MED-MVC and reverse osmosis (RO)), the necessary power is imported
in the majority of cases. If use is made of renewable energy sources, however, such
desalination processes can also be configured as extensively self-supplying systems
where the power generated by, for example, photovoltaic systems or wind generators
is used directly for desalination.
With regard to the various power supply configurations, a fundamental distinction
must be made between:

• Standalone single-purpose plants that produce desalinated water only (thermal


process or membrane process)
• Dual-purpose plants that produce both power and water (seawater desalination
and power plant) In such cases, the desalination part can use:
– A thermal process (MSF, MED, or MED-TVC) (Fig. 2.23).
– Reverse osmosis (RO). In this case, a dual-purpose plant configuration refers
to the direct connection of the membrane process to a power plant with
partially shared use of infrastructure systems.
– A thermal process (MSF, MED, or MED-TVC) in combination with a power-
consuming system (reverse osmosis (RO) or MED-MVC) in a hybrid
configuration.
2.4 Seawater Desalination Processes and Their Development 57

Seawater
intake Seawater feed

Electricity
Steam

Brine, Concentrate
Generator
Cooling water
Turbine
Power export Product

Carbon dioxide

Condensate

Fuel Steam
Generator

Product
Thermal Desalinaon Potabilizaon

Ouall

Fig. 2.23 Configuration of dual-purpose plant

Most single-purpose plants use reverse osmosis as desalination. Where seawater


conditions are difficult or where surplus heat is available from power plants or
industrial processes, the MED or MED-TVC thermal process is also used for
single-purpose desalination. Because of its lower unit capacity (see Table 2.7),
however, the use of MED-MVC is limited mainly to applications with lower plant
capacities.
MSF has been the predominant process for seawater desalination and power
plants (dual-purpose plants). However, the MED-TVC process is now also being
used to an increasing extent for the desalination part of such systems.
The first hybrid seawater desalination and power plants employed a combination
of MSF with RO for the desalination process. Nowadays, however, MED-TVC is
also being increasingly used as a thermal process in hybrid plants.
58 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

2.4.2 Development of Reverse Osmosis as a Seawater Desalination


Process

In terms of its commercial and technological development, seawater desalination has


in recent decades received a significant stimulus from GCC countries in the Persian
Gulf. Although desalination technology continues to benefit from this stimulus, there
is also a growing influence from non-GCC parts of the world.
Beginning in the 1970s, significant growth in population, economic activity,
and quality of life in the countries of the Persian Gulf, thanks to their wealth of oil
and natural gas, has led in those countries to a constantly rising demand for water and
electricity. Because of the shortage of natural water resources in these arid regions,
the gap between water demand and available freshwater from natural resources made
it necessary to develop seawater desalination technology as a large-scale source of
freshwater and to make suitable process capacities available.
In the 1960s and 1970s, the two thermal processes of multi-stage flash evapora-
tion (MSF) and multi-effect distillation (MED) were commercially available as
suitable processes. In the Gulf region, especially in Kuwait, various smaller-scale
single-purpose plants with unit capacities of between 4500 and 9100 m3/day (1–2
MIGD) had been in operation since the 1950s and 1960s (Shuwaik, Shuaiba)
[22]. At that time, MED plants were in worldwide operation, mostly in the industrial
sector, with similar unit capacities (St. Croix, Virgin Islands). The reverse osmosis
process was not yet available on an industrial scale, nor had it attained the opera-
tional maturity or reliability necessary to construct plants with capacities suitable for
integration into the supply systems of the Gulf. Also, as far as the two thermal
processes were concerned, operational experience with MSF in existing low- and
medium-capacity plants was more positive than for MED. The MSF process showed
itself to be significantly less sensitive than the MED process to scaling and fouling or
generally to variations in seawater quality.
In the 1970s, the first dual-purpose plants, in which the MSF desalination process
was integrated into a power plant, went into operation in Saudi Arabia (Jeddah I, II,
and III) [23]. Such a seawater desalination and power plant is capable of producing
both water and power. The process heat required for the thermal process is extracted
in the form of steam from the low-pressure stages of condensing turbines or is drawn
from the output stage of a backpressure turbine. Subsequently, more and larger dual-
purpose plants were constructed in rapid succession at the end of the 1970s and in the
1980s not only in Saudi Arabia but also in other GCC countries, such as Kuwait, the
United Arab Emirates, Qatar, and Oman. In the course of this around 10-year period,
the unit capacity of MSF plants grew from initially 11,400 m3/day (2.5 MIGD) to
28,000 m3/day (6.2 MIGD). The installed water capacity of these desalination and
power plants, which was originally around 23,000 m3/day (5 MIGD), had soared by
the end of the 1980s to approximately 1200,000 m3/day (250 MIGD) at the Al Jubail
desalination complex in Saudi Arabia. Also in the early to mid-1990s, the growth of
combined power and water supply systems was sustained in the Gulf states in the
form of dual-purpose plants, albeit on a reduced scale with a smaller number of new
plants in comparison with the 1970s and 1980s. During this period, however, MSF
2.4 Seawater Desalination Processes and Their Development 59

unit capacity grew to 56,800 m3/day (12.5 MIGD). The installed capacities of the
new plants were in the range between 27,000 m3/day (5.9 MIGD) (Al Gubrah,
Oman) and 341,000 m3/day (75 MIGD) (Al Taweelah B2, Abu Dhabi, UAE).
In the period from 1970 to 1995, globally installed seawater desalination capacity
and its growth were substantially determined by the construction and extension of
dual-purpose plants in the Gulf states (see Fig. 2.24).
High requirements in terms of security of supply of integrated systems in the Gulf
states were satisfactorily met by the MSF process with its operational robustness and
insensitivity to variations in raw water quality. Consequently, MSF was for decades
the process of choice for the desalination part of dual-purpose plants and thus the
predominant seawater desalination process not only in the GCC countries but also
worldwide.
In non-GCC countries, seawater desalination was used mainly in single-purpose
applications in industry and in power plants for process water supply and drinking
water production for smaller municipalities. This was also the principal area of
application for the second thermal process, multi-effect distillation (MED). Also
with this process, unit capacity gradually increased from 4500 to 9000 m3/day (1–2
MIGD) in the 1960s to 22,700 m3/day (5 MIGD) of MED-TVC systems by the end
of 2000 (1999, Layyah, Sharjah, UAE). The goal was to make the process competi-
tive with MSF also for dual-purpose plants. MED-TVC was used for the desalination
part of a dual-purpose plant for the first time in 2002, when an extension was built to
the Taweelah seawater desalination complex, the Taweelah A1 plant. The plant’s
desalination capacity of 242,000 m3/day is provided by 14 MED units, each with a
unit capacity of 17,275 m3/day (3.8 MIGD).

35
Cumulative
online
desalination 30 World
capacity
[x106 m3/d] Gulf
25
Non - Gulf

20

15

10

0
1970 1980 1990 2000 2010 2020
Year

Fig. 2.24 Development of seawater desalination capacity, world and Gulf (data source: Global
Water Intelligence (GWI) DesalData)
60 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

MED-TVC unit capacity continued to increase in the following years from


around 29,000 m3/day (6.3 MIGD) and 39,000 m3/day (8.5 MIGD) up to
68,000 m3/day (15 MIGD) (see Table 2.7). This growth in unit capacity, combined
with lower power consumption in comparison with MSF, is leading increasingly to
the use of MED in modern dual-purpose plants. However, the unit capacity of MSF
plants has also risen. The unit water capacity of the MSF process in 2013 reached
110,000 m3/day (24 MIGD).
The integration of thermal seawater desalination into the heat cycle of a power
plant is characterized, depending on the type and configuration of power plant, by a
certain relationship between generated electrical power and water production capac-
ity. This relationship is determined primarily by the power/heat ratio of the power
plant. The quantity of water that can be produced from the heat made available by the
power plant is a function of the efficiency of heat utilization in the thermal desalina-
tion process. The higher the performance ratio of the thermal process, the more water
can be produced in relation to the power generated by the power plant. This
relationship forms part of the specific design of a dual-purpose seawater desalination
and power plant, and, to avoid unfavourable impacts on the operating economics of
the power plant, it can be varied only within a relatively narrow range. Particularly in
the Gulf countries, however, the consumption of power and water varies greatly
depending on the time of year, with demand for both power and water peaking
during the summer months, but with significantly reduced power consumption in
winter due to substantially less use of air-conditioning. Water consumption, how-
ever, does not fall in winter to the same extent as power consumption. Figure 2.25
illustrates this situation with reference to load variation/load duration curves for

Fig. 2.25 Load variation/load duration curves for Abu Dhabi seawater desalination and power
plants (source: ADWEA, Abu Dhabi Water and Electricity Authority)
2.4 Seawater Desalination Processes and Their Development 61

power and water for seawater desalination and power plants in Abu Dhabi. A
consequence of this wintertime divergence between water consumption and power
consumption is that some of the power-generating capacity of the power plants is
taken out of service during the winter season and that, in order to maintain the water
production process, some of the heat for desalination is made available by auxiliary
boilers or by the existing plants without steam turbine operation.
For a significant proportion of the year, the dual-purpose plants are operated—
from the viewpoint of desalination—as single-purpose plants with considerably
higher energy consumption and reduced compensation of water production costs
by energy exports. To make dual-purpose plants more flexible with regard to their
water/power ratio, one option is to additionally equip them with desalination systems
that require only electrical power for their operation [24]. Reverse osmosis and
MED-MVC are such desalination processes that can be combined, in so-called
hybrid systems, with thermal processes. Owing to its limited unit capacity, however,
MED-MVC is not available for large seawater desalination and power plants.
In addition to the resulting greater flexibility of the desalination and power plants,
there is also improved energy efficiency on account of the lower specific energy
consumption of reverse osmosis as compared with thermal processes. Also, a hybrid
configuration provides cost advantages for the plant as a whole, thanks to the shared
use of infrastructure systems by power plant, thermal process, and reverse osmosis.
Variations in the product quality of the membrane process can be compensated by
mixing the distillate from the thermal process with permeate from reverse osmosis,
thereby making it possible to dispense with a second RO desalination stage
(Fig. 2.26).

Seawater feed
Seawater
intake Electricity

Heat

Brine, Concentrate
Cooling water

Power plant Product

Carbon dioxide

Thermal desalinaon Seawater reverse


process osmosis

Potabilisaon

Fig. 2.26 Dual-purpose hybrid configuration


62 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

The water capacity of dual-purpose plants in Saudi Arabia was expanded by


reverse osmosis as early as in 1989 and 1994 (Jeddah II, 113,600 m3/day, 25 MIGD).
This was followed by further capacity expansions to existing seawater desalination
and power plants by reverse osmosis in 1997 and 1998 for Medina-Yanbu with
128,000 m3/day (28 MIGD) and Al Jubail with an RO capacity of 90,900 m3/day
(20 MIGD). The first desalination and power plant to be designed from the outset as
a hybrid system, Fujairah I in the United Arab Emirates, with a water capacity of
454,600 m3/day (100 MIGD) (284,125 m3/day (62.5 MIGD) of which from MSF,
with 170,475 m3/day (37.5 MIGD) from reverse osmosis) went into service in 2008.
In the following decades, the hybrid configuration of dual-purpose plants has
found growing acceptance and application in the Gulf states for the expansion and
construction of desalination and power plants. The use of reverse osmosis in
dual-purpose plants in the Gulf has made a key contribution to broadening the
capacity spectrum of this membrane process from small- and medium-range systems
up to large-scale plants with product capacities of several hundred thousands of
cubic metres per day.
In the 1980s and particularly in the 1990s, reverse osmosis for seawater desalina-
tion gained increasingly in technical maturity and operational reliability in terms of
the available membranes and systems engineering. Especially the development of
efficient energy recovery systems led to constant improvements in the competitive-
ness of the process for use in single-purpose plants for freshwater supply to industrial
enterprises, power plants, and municipalities.
All such systems for reducing the energy consumption of seawater reverse
osmosis utilize the residual pressure still available in the RO concentrate to increase
the pressure of the seawater feed to the desalination plant. This residual pressure is
used mechanically either to increase the feed pressure or, after energy conversion, to
drive a high-pressure pump. The application of reverse-running centrifugal pumps or
so-called Francis turbines made it possible for the specific energy consumption
(SEC) of a seawater reverse osmosis stage to be lowered from over 8 kWh/m3 to
less than 6 kWh/m3. At the end of 1980, most of these systems were replaced by
Pelton turbines. This energy recovery system enabled the RO seawater desalination
stages of seawater desalination membrane plants to achieve specific energy
consumptions below 4–5 kWh/m3. SEC values below 4 kWh/m3 are also attainable
using turbochargers, which were employed for energy recovery from the early 1990s
onwards, initially in most cases in RO seawater desalination plants of lower plant
capacity. Driven by the RO concentrate, the turbocharger is a centrifugal pump that
acts as a so-called booster pump to generate the RO feed pressure in combination
with the high-pressure pump.
Up until the end of the 1990s, the Pelton turbine was the predominant energy
recovery system for RO. Then, at the beginning of 2000, isobaric energy
recovery systems, or so-called work exchangers, achieved technical maturity. In
the form of oscillating or rotating systems, these transfer the pressure of
the concentrate from the reverse osmosis stage directly to the seawater feed. This
allows the specific energy demand of a single-stage seawater reverse osmosis plant
to be lowered even further to below 3 kWh/m3.
2.4 Seawater Desalination Processes and Their Development 63

Fig. 2.27 Growth of different seawater desalination processes, Gulf and world (data source:
Global Water Intelligence (GWI) DesalData)

Such advances in mechanical engineering, membrane technology, and pre-


treatment efficiency gave rise, beginning with the 1990s, to rapid global growth of
reverse osmosis as a seawater desalination process and also to a swift increase in
installed plant capacity.
From the mid-1990s onwards, seawater desalination also began to record appre-
ciable growth outside of the Gulf states, especially as a result of the construction of
increasing numbers of large-capacity reverse osmosis plants in all parts of the world
(see Fig. 2.27).
64 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

The period from 2000 to 2012 was marked by a boom in seawater desalination.
From 1995 to 2012, the capacity of installed desalination plants rose by over
sixfold globally and by more than fivefold in the Gulf region, while other parts of
the world saw a more than tenfold increase. An annual comparison of newly
installed capacity for the three seawater desalination processes for this period
reveals that reverse osmosis recorded the highest global growth (Fig. 2.27). Even
though the MSF process also posted its hitherto highest growth rates in the Gulf
region over the same period, the MED process gained increasingly in significance
for dual-purpose plants, particularly as a hybrid process in combination with
reverse osmosis.
An especially striking fact about the reverse osmosis desalination process is its
rapid development from small- to medium-capacity to large-capacity plants
(Table 2.8). Whereas 46 plants with a capacity of over 10,000 m3/day and just
1 plant with a capacity of over 100,000 m3/day (Yanbu Medina) were newly
installed between 1991 and 2000, 270 plants with over 10,000 m3/day and
46 large-scale plants with over 100,000 m3/day were taken into service in the
boom period from 2006 to 2013. The largest plant capacities were 442,000 m3/day
(Hadera, Israel), 444,000 m3/day (Melbourne, Australia), 500,000 m3/day (Magtaa,
Algeria), and 624,000 m3/day (Sorek, Israel).
The installed SWRO capacity in non-Gulf countries increased at the same rate or
even faster than MSF capacity in the Gulf countries (Fig. 2.28).
At 2013, accounting for 46% of globally installed plant capacity, reverse
osmosis became the predominant seawater desalination process, followed by
MSF (42%) and MED (12%). In the Gulf region, the desalination parts of dual-
purpose plants however was still dominated by the MSF desalination process (64%
of installed plant capacity), albeit with a growing tendency to use MED (14%)
either on its own or in combination with reverse osmosis (22%) in hybrid systems
(Fig. 2.29).
For seawater reverse osmosis (SWRO), the annual growth in installed desalina-
tion capacity is continuing in the following years up to a share of this process of more
than 60% of the total desalination capacity of seawater in operation worldwide, with
tendency increasing also in the Gulf states to use reverse osmosis to extend the
capacity of hybrid plants and as well for installation of standalone desalination
systems.
Table 2.8 Growth of seawater reverse osmosis (SWRO) from 1970 to 2013 (data source: Global Water Intelligence (GWI) DesalData)
No. of plants installed worldwide during time
period Selected projects by capacity and significance
Time period >10,000 m3/day >100,000 m3/day Country Name Online date Capacity [m3/day] Type
1970–1985 12 – Saudi Arabia Jeddah 1978 12,000 SA
Malta Ghar Lapsi 1982 20,000 SA
1986–1990 9 – Saudi Arabia Jeddah I 1988 56,800 HS
Bahrain Al Dur 1990 45,500 SA
Spain Las Palmas 4 1990 12,000 SA
1991–1995 15 – Malta Pembroke 1991/1994 45,200 SA
Saudi Arabia Jeddah II 1994 56,800 HS
1996–2000 31 1 Japan Okinawa 1996/1997 60,000 SA
Spain Las Palmas III 1997 75,000 SA
Saudi Arabia Al Jubail 1997 90,909 HS
Cyprus Dhekelia 1997 40,000 SA
Saudi Arabia Yanbu-Medina 1998 127,800 HS
2001–2005 57 4 Spain Bahia de Palma 2001 68,000 SA
Cyprus Larnaca 2001 54.000 SA
2.4 Seawater Desalination Processes and Their Development

Spain Carboneras 2002 120,000 SA


Trinidad and Tobago Point Lisas 2002 109,100 SA
Israel Ashkelon 2005 326,100 SA
Singapore Tuas 2005 136,000 SA
2006–2013 270 46 USA Tampa Bay 2003/2007 108,830 SA
Chile Antofogasta 2006 52,000 SA
China Daishan 2007 108,000 SA
Australia Kwinana/Perth 2007 143,000 SA
Algeria El Hamma 2008 200,000 SA
UAE Fujairah I 2008 170,000 HS
65

(continued)
Table 2.8 (continued)
66

No. of plants installed worldwide during time


period Selected projects by capacity and significance
2

Time period >10,000 m3/day >100,000 m3/day Country Name Online date Capacity [m3/day] Type
Saudi Arabia Rabigh IWSPP 2008 218,000 SA
Oman Al Barka 2 2009 120,000 HS
Spain Barcelona 2009 200,000 SA
Oman Sur IWP 2009 80,200 SA
UAE Fujairah II 2010 136,500 HS
Israel Hadera 2010 442,000 SA
Saudi Arabia Shuqaiq 2 2010 213,000 HS
Australia Kurnell/Sydney 2010 250,000 SA
Bahrain Al Dur IWPP 2011 218,200 HS
Kuwait Shuwaikh 2011 136,400 HS
Spain Torrevieja 2011 240,000 SA
Algeria Magtaa 2012 500,000 SA
India Chennai 2012 100,000 SA
Saudi Arabia Jeddah III 2013 240,000 HS
Australia Melbourne 2013 444,000 SA
Israel Sorek 2013 624,000 SA
SA ¼ standalone HS ¼ hybrid dual purpose system
Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .
2.4 Seawater Desalination Processes and Their Development 67

Fig. 2.28 Growth of cumulative online seawater desalination process capacity, 1970–2012 (data
source: Global Water Intelligence (GWI) DesalData)

Fig. 2.29 Market shares of


seawater desalination
processes in 2013, Gulf and SWRO
worldwide 22%

MED MSF
14% 64%

Gulf

SWRO MSF
46% 42%

MED
12%

World
68 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

2.4.3 Potential and Future Prospects for Reverse Osmosis


in Seawater Desalination

The centralized co-production of power and water in dual-purpose plants, which is


typical of the GCC countries, however does not find identical acceptance or applica-
tion in the seawater desalination sector in most other parts of the world. As is
indicated by the growth of globally installed desalination systems, the trend in
non-GCC countries is towards single-purpose plants. This is a major advantage for
desalination plants using reverse osmosis technology. As a result the global trend
towards seawater desalination by reverse osmosis will continue and intensify as
demand increases for desalination plants not only in arid regions but also in densely
populated urban areas of the world.
The energy for the majority of seawater desalination plants comes from fossil
fuel-fired power plants. The abandonment of fossil fuels due to environmental
considerations as well as the increase in energy costs in parts of the world will
increasingly require the use of renewable energies such as solar, wind, and possibly
geothermal energy to supplement or replace existing power plants. As an
energy-intensive process, seawater desalination must already today address the
challenge of making the desalination process sustainable, i.e. climate-friendly.
This, too, will promote the use of renewable energy sources for seawater desalina-
tion. First steps have already been taken in Australia at reverse osmosis plants in
Perth, Sydney, and Melbourne, where the energy consumed by the plants is
generated by wind farms and fed into the conventional fossil-energy-based power
distribution network, from where the desalination plants draw their power.
Power-consuming desalination processes can be coupled with all variants of the
three above-mentioned renewable-energy-generating processes [25]. Thermal desa-
lination processes, however, must be directly connected to the heat-generating
systems, because heat transport over long distances is too expensive. Consequently,
solar thermal, i.e. CSP (concentrated solar power), is the most likely source of
energy for evaporation processes on an industrial scale (Fig. 2.30).
Also in this case, the coupling of desalination with power generation offers the
improved primary heat utilization known from fossil-fired plants. However, such
solar thermal power plants have a substantially bigger footprint than conventional
systems. Also, the radiation intensity at shoreline locations is often lower than at
inland locations. This may result, in some regions, in the existing coastal locations of
dual-purpose plants being unable to provide sufficient area for high-capacity CSP
seawater desalination and power plants. Moreover, higher inland radiation intensity
may make such locations more economically attractive for solar thermal power
plants than those near the sea. Generally, therefore, more intensive use of renewable
energy sources may further reinforce the trend towards the separation of power
generation and water production, i.e. towards seawater reverse osmosis.
Reverse osmosis for seawater desalination will therefore continue to play an
essential and increasingly important role as part of municipal and regional water
supply systems, in industrial applications, for power plants and for water supply in
the tourism sector.
2.4 Seawater Desalination Processes and Their Development 69

Fig. 2.30 Diagram of CSP power plant with MED


70 2 Earth’s Freshwater Resources and Their Management and the Role of Seawater. . .

References
1. World Water Assessment Programme (WWAP), “World Water Development Report WWDR1,
“Water for People, Water for Life”; Chapter 4-The Natural Water Cycle,” UNESCO, 2003.
2. World Water Assessment Programme (WWAP); UNESCO & WMO, with IAEA, “World
Water Development Report WWDR2: “Water, a Shared Responsibility”; Chapter 4 - The
State of the Resource,” UNESCO, 2006.
3. United Nations Environment Management Group., “Global Drylands: A UN system-wide
response,” United Nations, 2011.
4. UNCCD United Nations Convention to Combat Desertification, “Desertification - A visual
synthesis,” UNCCD, Zoï Environment Network, 2011.
5. Houerou, H. N. L., “Climate change, drought and desertification,” Journal of Arid
Environments, vol. 34, p. 133–185, 1996.
6. UNESCO-IHP International Hydrological Programme, Division of Water Sciences, “Atlas of
Transboundary Aquifers,” UNESCO, 2009.
7. UNESCO, “World Water Development Report WWDR3: "Water in a Changing World",”
UNESCO, 2009.
8. UNESCO-WWAP, “Climate Change And Water - An Overview from the World Water
Development Report 3: Water in a Changing World,” UNESCO, 2009.
9. Shiklomanov, I. A., “World Water Resources and their Use- a joint SHI/UNESCO product,”
UNESCO/IHP, St. Petersburg, 1999.
10. Sandra L. P., Daily G.C., Ehrlich P.R., “Human Appropriation of Renewable Fresh Water,”
Science, New Series, vol. 271, no. 5250, pp. 785-788, 9 February 1996.
11. Rost S., Gerten D., Bondeau A., Lucht W., Rohwer J., Schaphoff S., “Agricultural green and
blue water consumption and its influence on the global water system,” WATER RESOURCES
RESEARCH, vol. 44, no. W09405, doi:10.1029/2007WR006331, pp. 1 -17, 2008.
12. Mekonnen M.M., Hoekstra A.Y. , “National Water footprints accounts: The green, blue and
grey water footprint of Production and consumption,” UNESCO-IHE, Twente Water Centre,
University of Twente, Enschede, The Netherlands, 2011.
13. Morris B. L., Lawrence A. R. L., Chilton P.J. C, Adams B., Calow R. C., Klinck B. A.,
“Groundwater and its Susceptibility to Degradation: A Global Assessment of the Problem and
Options for Management,” United Nations Environment Programme, Nairobi, Kenya, 2003.
14. Alcamo J., Märker M., Flörke M., Vassolo S., “Water and Climate: A Global Perspective-The
Kassel World Water Series: Report 6,” 2003.
15. Alamco J., Henrichs T., Rösch T., “World Water in 2025: Global modeling and scenario
analysis for the World Commision on Water for the 21th Century; The Kassel World Water
Series, Report 2,” 2000.
16. Menzel L., Flörke M., Matovelle A., Alcamo J., “Impact of socio-economic development and
climate change on water resources and water stress,” in Proc. 1st Internat. Conf. on Adaptive
and Integrative Water Management (CAIWA 2007), Basel.
17. Pereira L.S., Cordery I., Iacovides I., “Coping with water scarcity,” UNESCO, Paris, 2002.
18. Global Water Partnership, “Perspectives Paper - Towards Integrated Urban Water Manage-
ment,” GWP Global Water Partnership, Stockholm, 2011.
19. Kayaga S., Smout I. (eds.), “Water Demand Management in the City of the Future,” WED,
Loughborough University, Leicestershire, 2011.
20. SWITCH Training Kit, “Integrated Urban Water Management in the City of the Future - Modul
3 - Water Supply - Exploring the Options,” ICLEI European Secretariat GmbH, Freiburg, 2011.
21. Tortajada C., “Water Management in Singapore,” Water Resources Development, vol. 22, no. 2,
p. 227–240, June 2006.
References 71

22. Al-Bahou M., Al-Rakaf Z., Zaki H., Ettouney H., “Desalination experience in Kuwait,”
Desalination, vol. 204, pp. 403-415, 2007.
23. Nada N., “Development of Desalination in Saudi Arabia,” in Water Arabia 2009, Manama,
Bahrain, 2009.
24. Ludwig H., “Hybrid systems in seawater desalination - practical design aspects, present status
and development perspectives,” Desalination, vol. 164, no. 1, pp. 1-18, 2004.
25. FICHTNER, DLR, Future Water, “MENA Regional Water Outlook - Part II - Desalination
Using Renewable Energy- Final Report,” World Bank, 2011.
Seawater: Composition and Properties
3

3.1 Composition and Temperature of Seawater

3.1.1 Composition

Seawater contains an abundance of organic and inorganic constituents. These are


present in dissolved, suspended, very finely distributed, or colloidal form.
The dissolved constituents of seawater can be assigned to the following main
categories:

• Inorganic main constituents that form the salt content of seawater with its cationic
and anionic components
• Inorganic substances that are contained in seawater only in small quantities or
traces, i.e. in the microgram range
• Dissolved gases, such as oxygen, nitrogen, and carbon dioxide
• Dissolved organic constituents that are produced by natural biological and chem-
ical processes (e.g. in the form of metabolites and decomposition products of
marine life), so-called natural organic matter (NOM)

Seawater also contains undissolved organic and inorganic substances in a


suspended, finely distributed or very finely distributed state. Organic matter is
present both in inanimate form and also as microbes and microorganisms (algae,
plankton, bacteria, etc.). Suspended matter is usually a mixture of organic and
inorganic substances and living organic substrates.
Especially in densely populated and heavily industrialized coastal regions, vari-
ous organic or inorganic substances can enter the sea after being discharged in the
effluent from industry and sewage treatment. This can affect the composition of
seawater, particularly with regard to its concentrations of dissolved heavy metals,
hydrocarbons, surfactants, and nutrients as well as its content of undissolved
substances. This results in a significant deterioration in the quality of the water.
Shipping along heavily trafficked sea routes as well as the offshore extraction of

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 73


H. Ludwig, Reverse Osmosis Seawater Desalination Volume 1,
https://doi.org/10.1007/978-3-030-81931-6_3
74 3 Seawater: Composition and Properties

crude oil and natural gas can also have local, yet increasingly extensive, adverse
impacts on seawater quality. A further problem is progressive pollution of the oceans
with plastic waste.
Both natural and human-caused organic substances, together with biological
activity in seawater, have a highly significant impact on the operation and availabil-
ity of technical processes that use seawater. This is particularly the case with
membrane processes for seawater desalination. For a detailed discussion of such
organic and biological fouling; the seawater constituents that cause it, how it is
influenced by external conditions, measures to protect technical systems against it,
and minimization of fouling, see [1] Chap. 2, in volume 2. The below remarks refer
exclusively to the inorganic constituents of seawater and their impacts on its physical
and physico-chemical properties.
All naturally occurring elements on Earth have been found in seawater, both in
high concentrations, such as in the main constituents of seawater salt content, and
also in very low and tiny concentrations in trace substances in seawater.
The salt content in the Earth’s seas and oceans differs very widely (see Table 3.1).
The salt content ranges from a very low 3–8 g/kg in the Baltic Sea to an average
32–38 g/kg in the Atlantic, Pacific, and Indian Oceans to maximum salinity values of
42 g/kg in the Red Sea and sometimes even up to 49 g/kg in the Arabian Gulf.
The salt contained in seawater consists mainly of six saline components, namely,
the cations sodium, magnesium, calcium, and potassium and the anions chloride and
sulphate. These saline components account for over 99% of the total salt content of
seawater (see Fig. 3.1). The remaining around 1% of salt content consists mainly of
the cations strontium and barium and the anions bicarbonate, carbonate, bromide,
borate, fluoride, and nitrate.
As far as the six main components of the salt contained in seawater are concerned,
the ratio of their respective concentrations in g of dissolved substance per kg of
seawater to the total seawater salt content in g/kg—yet also the ratio of their
concentrations with respect to each other—is broadly constant irrespective of sea-
water salt content. This principle of constant proportions—known also as
Forchhammer’s or Marcet’s principle after the oceanographers who first
investigated this natural phenomenon back in the nineteenth century and formalized
it as a law—makes it possible, with knowledge of the concentration of just one of the
main components of the salt content in a seawater sample, to estimate the total salt
content as well as the concentrations of the other main components.

Table 3.1 Salt content of Seawater surface salt content


seas and oceans on Earth
Sea or ocean g/kg
Baltic Sea ~ 3–8
Caspian Sea ~10–15
Pacific Ocean ~ 32–36
Atlantic Ocean ~34–37
Indian Ocean ~ 32–38
Mediterranean Sea ~38–40
Red Sea ~38–42
Arabian Gulf ~41–49
3.1 Composition and Temperature of Seawater 75

Based on this law of constant proportions, oceanographers have defined a stan-


dard seawater with a salinity of 35 g/kg and a defined ratio of the concentrations of
its main saline components, especially its chlorinity and conductivity (IAPSO
standard seawater) (see also Sect. 3.1 Chlorinity and Salinity). The seawater in
question is natural seawater that was taken from the North Atlantic and, having
been filtered and adjusted to the standard values, was used as a reference standard in
oceanography. With increasing accuracy in the analysis of the constituents of
seawater and in connection with the formulation of the “International Thermody-
namic Equation of Seawater—2010—TEOS 10”, this standard seawater was then
used to define a reference seawater (see Table 3.2), in which the sum of its
constituents is equivalent to a salt content of 35.165 g/kg [2].
Inorganic saline components of seawater not included among the above-
mentioned main components, especially those elements that occur in small or tiny
concentrations as well as compounds involved in biological and chemical processes,
can be subject to very significant local and seasonal variations.
The principle of constant proportions is at its most valid in the open sea. Near the
coast, particularly in shallow coastal waters, river estuaries, or sea bays, even the
main constituents of seawater are subject to anomalies, i.e. deviations from the
principle of constant proportions.

Fig. 3.1 Main components of seawater salt content


76 3 Seawater: Composition and Properties

Table 3.2 Composition of reference seawater


Ratio
Solute Solute per
Concentration per TDS chloride
mg/kgSW g-mol/
Component Symbol g/kgSW ppm kgSW % %
Calcium Ca++ 0.41208 412.08 0.010282 1.172 2.129
Magnesium Mg++ 1.28372 1283.72 0.052817 3.651 6.633
Sodium Na+ 10.78145 10,781.45 0.468983 30.660 55.710
Potassium K+ 0.3991 399.10 0.010208 1.135 2.062
Strontium Sr++ 0.00795 7.95 0.000091 0.023 0.041
Carbonatea CO3 0.01434 14.34 0.000239 0.041 0.074
Bicarbonatea HCO3 0.10481 104.81 0.001718 0.298 0.542
Sulphate SO4 2.71235 2712.35 0.028235 7.713 14.015
Chloride Cl 19.35271 19,352.71 0.545869 55.035 100.000
Bromide Br 0.06728 67.28 0.000842 0.191 0.348
Borona B(OH)3 0.01944 19.44 0.000314 0.055 0.100
[B444(OH)4] 0.00795 7.95 0.000101 0.023 0.041
Fluoride F 0.0013 1.30 0.000068 0.004 0.007
Hydroxyl OH 0.00014 0.14 0.000008 0.0004 0.001
Total ∑ 35.165 35,165 1.12
a
at pH ¼ 8.1

The water intake for seawater desalination plants comes mainly from coastal
areas. Annex 3.A1, Table 3.15 provides component analyses (full analyses) with
minimum and maximum salt contents and concentrations of individual components
for such coastal seawater intakes for various seas and oceans. The analysis
concentrations are given in the volume-specific unit mg/l, which is customary in
the field of seawater analysis. For comparison with the reference seawater analysis in
Table 3.2, these analysis data must be converted to mass-specific values (mg/kg) (see
also Sect. 3.1 Total Dissolved Solids (TDS)).

Chlorinity and Salinity


The salt content (S) of seawater is defined as the quantity of dissolved inorganic
substance per kilogram of seawater, i.e. the mass proportion of dissolved inorganic
substance. The salt content includes not only ionogenic but also neutral undissoci-
ated and weakly dissociated inorganic components. However, it does not include
dissolved organic substances.
Oceanographers refer to the salt content of seawater as its salinity. In terms of the
methodology of its determination, a distinction is made between different types of
salinity:
3.1 Composition and Temperature of Seawater 77

• Salinity (S), which is calculated from the chlorinity (Cl) of the seawater
• Practical salinity (SP), which is determined from the ratio of the conductivity of
the measured sample to the conductivity of standard seawater
• Reference salinity (SR), which corresponds to the salt content of standard
seawater as determined by analysis of the individual components
• Absolute salinity (SA), which is determined from the reference salinity while
taking into account the so-called salinity anomaly

For decades in the field of marine science, chlorinity was the parameter that
allowed a simple determination of the salinity (S) of seawater using Eq. (3.1). The
factor ( fCl St) yields the salinity (S) with reference to standard seawater with
SSt ¼ 35 g/kg and a chlorinity (Cl) of 19.374 g/kg. The previously used dimension

[ / ] for salinity (S), known also as ppt (parts per thousand), corresponds to the units
g/kg ¼ 1.000 mg/kg ¼ 1.000 ppm.
Using the factor ( fCl R), the reference salinity (SR), as defined in TEOS-10, is
calculated from the chlorinity as in Eq. (3.2):

S ¼ f Cl St  Cl ð3:1Þ

SR ¼ f Cl R  Cl ð3:2Þ


S ¼ salinity [ /  ], [ppt], [g/kg]
SR ¼ reference salinity [g/kg]
fCl St ¼ 1.80655—based on standard seawater
fCl R ¼ 1.815069—based on reference seawater
Cl ¼ chlorinity [ / ], [ppt], [g/kg]

Chlorinity is measured by Knudsen-Mohr titration of a seawater sample with


silver nitrate solution. This determines all the halogenides present in the sample,
i.e. not only chloride but also the concentrations of bromide and iodide (of which
there is a negligibly small concentration in seawater):

Agþ þ seawater ! AgClðSÞ þ AgBrðsÞ þ AgJðsÞ

Therefore, the chlorinity of a seawater is equivalent to the sum of its chloride and
bromide concentrations.
With knowledge of the ratio of the chlorinity of a seawater to that of standard
seawater, it is possible to determine the concentrations of the other main components
of the seawater. For this purpose, a concentration factor fCF can be calculated as in
Eq. (3.3). By multiplying the concentration of the wanted component in the standard
or reference seawater by this concentration factor, one obtains the wanted concen-
tration of the particular salt content component in the seawater sample as in Eq. (3.4).
This makes it possible to use the principle of constant proportions, together with the
standard or reference seawater, to check the validity of a seawater analysis:
78 3 Seawater: Composition and Properties

ClS
f CF ¼ ð3:3Þ
ClSt
cCs ¼ cCSt  f CF ð3:4Þ

fCF ¼ concentration factor


ClS ¼ chlorinity of sample [g/kg], [mg/kg]
ClSt ¼ chlorinity of standard [g/kg], [mg/kg]
cCS ¼ concentration of component in sample [g/kg], [mg/kg]
cCSt ¼ concentration of component in standard seawater [g/kg], [mg/kg]

With the introduction of the Practical Salinity Scale (PSS-78) standard [3–5],
chlorinity was replaced by conductivity as the standard method of measuring and
calculating the salinity of a seawater.
Practical salinity, which was introduced with the PSS-78 standard as the mea-
sure of the salinity of seawater, is calculated from the ratio K15 of the electrical
conductivity of a seawater sample at a temperature of 15  C and the pressure of a
standard atmosphere and the conductivity value of a potassium chloride (KCl)
solution with a mass fraction of 32.4356 g/kg at identical temperature and pressure.
The conductivity of the potassium chloride standard of 42.914 mS/cm corresponds
to the conductivity value of standard seawater (see Eqs. 3.5 and 3.6):
   
C Sp , t ¼ 15 C, 0 C Sp , t ¼ 15 C, 0
K 15 ¼ ¼ ð3:5Þ
Cð35, t ¼ 15 C, 0Þ 42:914

K15 ¼ conductivity ratio at standard conditions


C(SP, t ¼ 15  C, 0) ¼ conductivity of seawater sample of unknown salinity,
temperature of 15  C, and atmospheric pressure
C(35, t ¼ 15  C,0) ¼ conductivity of KCl standard equal to Sp ¼ 35, temperature of
15  C, and atmospheric pressure; CStandard ¼ 42.914 mS/cm

1=2 3=2 5=2


Sp ¼ a0 þ a1 K 15 þ a2 K 15 þ a3 K 15 þ a4 K 215 þ a5 K 15 ð3:6Þ

a0 ¼ +0.008 a3 ¼ +14.0941
a1 ¼ 0.1692 a4 ¼ 7.0261
a2 ¼ +25.3851 a5 ¼ +2.7081
SP ¼ practical salinity.

If ratio K15 ¼ 1, then, by definition, practical salinity Sp will have a value of


SP ¼ 35, i.e. the salinity of standard seawater.
If sampled in situ, however, the conductivity of a seawater sample is measured at
the prevailing water temperature, in which case the conductivity ratio R is obtained
3.1 Composition and Temperature of Seawater 79

using Eq. (3.7). As in Eq. (3.8), R is composed of the in situ conductivity ratio Rt and
the factors for pressure correction Rp and temperature correction for the standard KCl
solution rt:
   
C Sp , t, 0 C Sp , t, 0
R¼ ¼ ð3:7Þ
C ð35, t ¼ 15 C, 0Þ 42:914

R ¼ conductivity ratio
C(SP, t,0) ¼ conductivity of seawater sample of unknown salinity, in situ tempera-
ture and atmospheric pressure

R ¼ Rp  Rt  r t ð3:8Þ

Rp ¼ pressure correction factor ¼ 1 for atmospheric conditions


Rt ¼ conductivity ratio at in situ conditions
rt ¼ temperature correction factor for standard KCl solution

The dependence of practical salinity on pressure is negligible where the water is


drawn from the surface or a shallow depth (less than 100 m), as is the case for a
seawater desalination plant.
Rt is then calculated from Eqs. (3.9 and 3.10) and SP is determined as in
Eq. (3.11). The values for Sp obtained from Eqs. (3.6 and 3.11) have no physical
dimension, as they were calculated from a ratio of conductivities. However, the
values for practical salinity are often given in psu (practical salinity unit):

R
Rt ¼ ð3:9Þ
rt

r t ¼ c0 þ c1  t þ c2  t 2 þ c3  t 3 þ c4  t 4 ð3:10Þ

c0 ¼ +0.6766097 c3 ¼  6.969810–7
c1 ¼ +2.0056410–2 c4 ¼ +1.003110–9
c2 ¼ +1.10425910–4

t ¼ in situ temperature during conductivity detection

1=2 3=2 5=2


Sp ¼ a 0 þ a1 R t þ a2 Rt þ a3 Rt þ a4 R2t þ a5 Rt þ ΔS ð3:11Þ

ðt  15Þ  1=2 3=2 5=2



ΔS ¼ b0 þ b1 Rt þ b2 Rt þ b3 Rt þ b4 R2t þ b5 Rt
1 þ kðt  15Þ
80 3 Seawater: Composition and Properties

b0 ¼ +0.0005 b4 ¼ +0.0636
b1 ¼ 0.0056 b5 ¼ 0.0144
b2 ¼ 0.0066 k ¼ +0.0162
b3 ¼ 0.0375

Equations (3.6 and 3.11) can be used according to [5] to calculate practical
salinity in a range of seawater salt content of 2–42 psu and in a temperature range
of 2–35  C. For technical measurements, however, the range of validity of salinity
can be extended to Sp ¼50 without significant loss of the required accuracy. For
desalination plant concentrates of higher salinity, the practical salinity can be
determined if the concentrates are diluted to the range of validity of the PSS-78
standard.
The temperature dependence of practical salinity according to the equations of the
PSS-78 standard is based on the International Practical Temperature Scale 1968
(IPTS-68). In 1990, this scale was superseded by the International Practical Tem-
perature Scale 1990 (IPTS-90). Equation (3.12) is used to convert temperatures
according to IPTS-90 to temperatures according to IPTS-68, i.e. application of
Eqs. (3.10 and 3.11) with measured in situ temperatures.

t 68 ¼ 1:00024  t 90 ð3:12Þ

t68 ¼ temperature according to IPTS-68 [ C]


t90 ¼ temperature according to IPTS-90 [ C]

As temperature correction of the temperature input value results in changes only


in the third or fourth decimal place, this adjustment, which is necessary for scientific
purposes, can be neglected for engineering calculations in connection with seawater
desalination. In the following, therefore, no distinction is made between IPTS-68
and IPTS-90 where a temperature is given in an equation.
With the introduction of TES-10 and the associated calculation of the physical
properties of seawater on a thermodynamic basis, dimensionless practical salinity SP
could no longer be used as an input value for TES-10 calculation algorithms. This
led to the SR being given as the mass concentration g/kg in SI units [2].
Determination of the reference salinity is based on its calculation from the sum of
the individual ion concentrations in standard seawater. Determination of the
concentrations of the anions bicarbonate HCO3, carbonate CO3, and borate B
(OH) 4 is based on a pH value of seawater of 8.1. The thus obtained value for
reference salinity is 35.165 g/kg, which is currently the most accurate value for the
actual salt content of IAPSO standard seawater (see Table 3.2).
SR is calculated as in Eq. (3.2) from chlorinity and from SP as in Eq. (3.13):
3.1 Composition and Temperature of Seawater 81

35:165
SR ¼  SP ¼ 1:004715  SP ð3:13Þ
35

SR ¼ reference salinity [g/kg]

The conductivity of seawater is measured mainly with reference to the heavily


dissociated cations and anions of the salt content, i.e. those components that are
determining for the principle of constant proportions. On the other hand, weakly
dissociated and undissociated dissolved inorganic constituents, such as nitrates,
phosphates, and silicates, are more or less left out of consideration. Equally
neglected are changes in lime/carbonic acid equilibrium as a result of different
carbon dioxide solubility through differences in pH value and temperature of the
seawater as well as pH value-dependent changes in the ratio of boric acid to borate.
SA takes account of the errors resulting from the measurement of conductivity.
This is done by providing the reference salinity with a correction value, the so-called
salinity anomaly (δSA). SA is then calculated from SR as follows:

SA ¼ SR þ δSA ð3:14Þ

SA ¼ absolute salinity [g/kg]


δSA ¼ salinity anomaly [g/kg]

In the open ocean, the values of the salinity anomaly are assigned to specific
latitudes and longitudes, and the TEOS-10 algorithm for calculating SA contains a
database with corresponding values, which are kept up to date by continuous
measurement of the anomalies. This method, however, cannot be used in coastal
waters, where δSA must be individually determined on a case-by-case basis [6].
The amount of required correction of the reference salinity using the salinity
anomaly is a maximum 0.03 g/kg in the open ocean. In coastal waters, the anomaly
can be up to 0.1 g/kg. For salinity values of 30–50 g/kg in the seas and oceans of the
Earth, therefore, correction of the salt content caused by the salinity anomaly is less
than 0.5%.
For technical application of conductivity measurements in seawater desalination,
therefore, absolute salinity can be equated with reference salinity Eq. (3.15):

SA ffi SR ffi 1:004715  SP ð3:15Þ

For online measurement of the seawater intake into a desalination plant or to


determine the gradient of salt content increase when the concentrate from a seawater
desalination plant is discharged into the ocean, it is readily possible, given the
required accuracy of the results, for SR or even SP to be used to evaluate the
measurement results. To calculate the physical properties of seawater for technical
purposes, however, it is advisable to use SR or, as in Eq. (3.15), SA, not SP.
82 3 Seawater: Composition and Properties

Fig. 3.2 Dependence of seawater conductivity on SA and temperature

The graph in Fig. 3.2 presents the dependence of seawater conductivity on water
temperature in the range 10–35  C and absolute salinity in the range 2–50 g/kg,
based on Eqs. (3.7, 3.9, 3.10, 3.11, 3.13, and 3.14).

Total Dissolved Solids (TDS)


In seawater desalination technology, determination of the salt content of seawater is
based not on salinity, but on TDS (total dissolved solids) content. According to
ASTM D4195-08 Standard Guide for Water Analysis for Reverse Osmosis and
Nanofiltration Application, the TDS content is determined by analysis of the
dissolved inorganic components of seawater, i.e. as the sum of the cation and
anion concentrations of those components (Table 3.3, Eq. 3.16). The standard lists
the most important components for analysis and the relevant ASTM analysis
standards [7]:

X
i¼6 X
i¼13 X
i¼15
TDSV ¼ CCi þ C Ai þ C Oi ð3:16Þ
i¼1 i¼7 i¼14
3.1 Composition and Temperature of Seawater 83

Table 3.3 Cations and anions for seawater analysis, their molecular/atomic weights, and valencies

Cations and anions for analysis of seawater Atomic/molecular weight Mi Valency Zi


components g/mole –
Cations Ci
1 Sodium Na+ 22.990 1
2 Potassium K+ 39.098 1
3 Calcium Ca++ 40.078 2
4 Magnesium Mg++ 24.305 2
5 Strontium Sr++ 87.620 2
6 Barium Ba++ 137.327 2
Anions Ai
7 Chloride Cl 35.453 1
9 Nitrate NO3 62.0049 1
8 Sulphate SO4 96.0626 2
10 Bromide Br 70.904 1
10 Bicarbonate HCO3 61.0168 1
11 Carbonate CO3 60.008 2
12 Fluoride F 18.998 1
13 Borate [B(OH)4] 78.8404 1
Other components Oi
14 Boric acid B(OH)3 61.833
15 Silica SiO2 60.084

TDSV ¼ total dissolved solids content, volume-based [mg/l, g/l, g/m3]


C Ci ¼ concentration of cation component I [mg/l, g/m3]
C Ai ¼ concentration of anion component I [mg/l, g/m3]

According to the customary units of measurement in seawater analysis, the


concentrations of the individual components, and therefore also the TDS, are
given as volume concentrations in mg/l or g/m3. The TDS in volume concentration
is referred to in the following as TDSV in order to distinguish it from TDSM in mass
concentration with the unit mg/kg, ppm, or g/kg.
The cations, anions, and dissolved neutral inorganic components listed in
Table 3.3 are the seawater components necessary for the design of a reverse osmosis
membrane plant, for determination of the probable permeate quality and for calcula-
tion of the scaling potential. Additional inorganic components, such as heavy metals,
are normally also determined in order to obtain information on the fouling potential
of the seawater. These, however, are present in seawater in such low concentrations
that they have no appreciable influence on the TDS value of the seawater.
Experience shows that the accuracy of analysis in determining the principal
components sodium and chloride in seawater is subject to wide variation (depending
on the analysis method and experience of the analysis laboratory in seawater
analysis) [8]. Therefore, the results of seawater analyses should always be verified
for consistency and plausibility. Such quality control involves performing an
84 3 Seawater: Composition and Properties

electroneutrality test with the analysis. According to the principle of


electroneutrality, the sum of positive cation equivalents in a solution must be
equal to the sum of negative anion equivalents. The cation and anion equivalents
∑Ceq, ∑Aeq of an analysis are calculated from the equivalents of their individual
components as follows:

X X
i¼6
C ci  Z i
C eq ¼ ð3:17Þ
i¼1
Mi

X X
i¼13
C Ai  Z i
Aeq ¼ ð3:18Þ
i¼7
Mi

∑Ceq ¼ sum of cation equivalents [meq/l, geq/m3]


∑Aeq ¼ sum of anion equivalents [meq/l, geq/m3]

The deviation Δeq from electroneutrality in an analysis, i.e. the analysis error, is
determined from Eq. (3.19). In seawater analysis, Δeq, % as in Eq. (3.20) should not
exceed a value of 5% and, for good analysis accuracy, should be in the range of
2–3%:
P P
C eq  Aeq
Δeq ¼ P P ð3:19Þ
C eq þ Aeq
 
△eq,% ¼ Δeq   100 ð3:20Þ

Δeq ¼ deviation from electroneutrality


Δeq,% ¼ percentage deviation from electroneutrality [%]

To use the analysis results for the design or operational monitoring of a seawater
desalination plant, quality control is followed by an electroneutrality adjustment. For
this purpose, depending on whether the balance shows an increased concentration of
cations or anions (i.e. whether the difference of cation and anion equivalents is
positive or negative), the analysis is balanced according to the difference of ion
equivalents through the addition of cations (usually sodium Na) or anions (usually
chloride Cl) Eqs. (3.21 and 3.22):
For positive Δeq:

M Na
ΔNa ¼ Δeq  ¼ Δeq  22:99 ð3:21Þ
Z Na
For negative Δeq:
3.1 Composition and Temperature of Seawater 85

M Cl
ΔCl ¼ Δeq  ¼ Δeq  35:453 ð3:22Þ
Z Cl

ΔNa ¼ increase of Na concentration for balance of analysis


ΔCl ¼ increase of Cl concentration for balance of analysis

The above-described seawater analysis includes all inorganic components,


whether heavily dissociated, weakly dissociated, or undissociated. Consequently,
the thus determined TDS (total dissolved solids) content corresponds to SA
Eq. (3.23):

TDSM ffi SA ð3:23Þ

This, however, applies only to TDSM in mass concentration units, because the
absolute salinity must be available as a mass concentration to allow it to be used as
an input value in the relevant algorithms, such as for calculating the physical
properties of seawater. To convert volume-based TDSV, which is more customary
in seawater analysis, to mass-based TDSM, TDSV must be divided by the density ρSW
of the seawater sample at the analysis temperature as follows:

TDSV
TDSM ¼ ð3:24Þ
ρSW,t

TDSM ¼ total dissolved solids content, mass-based [mg/kg, g/kg]


TDSV ¼ total dissolved solids content, volume-based [mg/l, g/l]
ρSW,t ¼ density of seawater sample at analysis temperature [kg/l]

Since, however, calculation of the density of the seawater sample requires


knowledge of TDSM, an exact solution of the problem is possible only by an iterative
calculation in which TDSV is initially used as a first approximation to determine the
density. More simply, TDSM can be calculated using the graph presented in Fig. 3.3.
The graph shows a correction factor Δ TDSV,20 C for a temperature of 20  C and a
TDSV range of 20–80 g/l, which can be used to calculate TDSM as follows:

TDSM ¼ TDSV  △TDS,20 C ð3:25Þ

Although this yields approximate values, these are of suitable accuracy for
technical calculations. Also in the temperature range of 15–25  C, this correction
factor can be used without significant loss of accuracy. With the thus obtained value
for TDSM, the density of the solution can then be determined according to 3.2.2.1
Eqs. (3.39 and 3.40). With knowledge of the density, it is possible, as with Eq. (3.24)
for TDSV, to convert the other individual components of the analysis from volume-
based to mass-based concentrations.
86 3 Seawater: Composition and Properties

Fig. 3.3 TDS correction factor for calculation of TDSM from TDSV

To check the plausibility of the seawater analysis, it is then possible, as described


above, to apply the principle of constant proportions and to compare the analysis
with standard seawater using chlorinity as in Eqs. (3.3 and 3.4).
TDS refers also to the salt content determined by concentrating a defined volume
of a filtered seawater sample, drying it at 180  C and weighing the residue. As this is
considerably easier and less labour-intensive than determining TDSV or TDSM by
component analysis, this method of salt content determination is often preferred for
verifying the performance and monitoring the operation of a desalination plant.
However, experience shows that, compared with calculation of TDS by component
analysis, this analysis method yields significantly less accurate, widely varying
values [8].
Possible reasons for an increased amount of residue are:

• Crystallization water can be left in salt components despite a drying temperature


of 180  C.
• The highly hygroscopic calcium chloride contained in the residue can attract
moisture from the air and bind it as crystallization water.
3.1 Composition and Temperature of Seawater 87

• In addition to inorganic salt components, the residue additionally contains


organic and finely colloidal substances.

The amount of residue can be reduced by thermal decomposition of bicarbonate


and loss of other volatile salt components, such as boric acid.
Determination of TDS by the concentration method produces highly inaccurate
and poorly reproducible values for seawater salt content. Consequently, results
obtained by this method should not be used to determine the physical and
physico-chemical properties of seawater or as a basis for the design and performance
monitoring of a desalination plant.

3.1.2 Temperature

In the near-surface area of the sea, from where most seawater is drawn for technical
use in desalination or as cooling water, apart from salinity it is mainly temperature
that has a significant influence on the properties of seawater. This applies to both
physical parameters and also physico-chemical behaviours, i.e. the course of chemi-
cal reactions and their time to equilibrium. However, also the biological processes in
the sea, especially in nutrient-rich coastal areas, are significantly influenced by water
temperature (e.g. algae and jellyfish bloom).
The range of average annual surface temperatures in the Earth’s seas and oceans
extends from values around freezing point at high northerly and southerly latitudes
to temperatures close to 30  C in tropical sea areas along the equatorial belt
(Fig. 3.4).

Fig. 3.4 Global annual average seawater surface temperatures (Source: National Oceanic and
Atmospheric Administration (NOAA)—National Centers for Environmental Information—Live
Access Server)
88 3 Seawater: Composition and Properties

Fig. 3.5 Typical ocean temperature depth profile

The water temperature in the Arabian Gulf in summer can reach values up to
37  C or higher (Annex 3.A1, Table 3.15 and Table 3.16).
Seawater temperature falls with increasing sea depth, albeit very differently at
low latitudes, mid-latitudes, and high latitudes (Fig. 3.5).
From the sea surface down to the deep sea, there are four defined zones with
different temperature behaviours:
• A mixing zone between the surface and down to around 20–60 m depth with
strong influencing by wind and heat exchange processes
• A seasonal thermocline down to around 100 m depth, which is subject to seasonal
temperature influences
• A permanent thermocline in the region between 800 and 1500 m, which is
influenced in particular by current flows
• The deep sea, where there is hardly any change in temperature

Although the surface temperature of the sea is higher in the equatorial zone, it
quickly falls with increasing depth. In the northern and southern seas in zones of
high latitude with very much lower surface temperature, the change in temperature
with increasing water depth is significantly smaller.
The temperature profiles presented in Fig. 3.5, however, are more characteristic of
the open sea. In the coastal areas of seas and oceans, which lie predominantly in the
mixing zone and the seasonal thermocline, the temperature profiles are very much
more complex, because, above all, seasonal influences (intensity of solar radiation,
air and land mass temperatures, surf intensity) have a very much stronger influence
on water temperatures there than in the open sea. Figure 3.6 presents the seasonal
changes in temperature depth profile down to a sea depth of 80 m for a location in the
western Mediterranean. In winter, there is hardly any difference between surface
temperature and depth temperature. In spring and summer, respectively, the differ-
ence in water temperature between surface and 80 m depth rises to approx. 2  C and
8  C. This phenomenon is reversed in the autumn.
3.1 Composition and Temperature of Seawater 89

Seawater temperature [°C]


10 15 20 25
0
-10
-20
-30
Depth [m]
-40
Winter
-50 Spring
Summer
-60 Autumn
-70
-80

Fig. 3.6 Seasonal temperature depth profile (Data source: Statistical Institute of Catalonia—
seawater, average temperature, at different depths)

Changes in temperature not only influence the above-described properties of


seawater but also have a considerable effect on the operating performance of
desalination plants and cooling systems. Thus, the energy demand of a seawater
desalination plant is heavily dependent on the water temperature at its intake.
Consequently, the design of such a plant must be based on the likely intake
temperature range (according to the water intake depth) and not on the temperatures
measured at the surface of the water.
The influence of water depth on seawater temperature is compounded by seasonal
changes in water temperature, which can vary very greatly depending on location
(Fig. 3.7).
With temperatures between a minimum of 18  C and maximum of 37  C,
seasonal temperature variations in the Arabian Gulf can be up to approx. 20  C.
There are also non-seasonal temperature changes lasting 1 or more days, which can
likewise be up to 5  C. In a tropical sea nearer the Equator, the seasonal influences
and short-time variations in seawater temperature are significantly smaller. As
demonstrated by the example of Singapore in Fig. 3.7, the seawater temperature
there varies seasonally by around 6  C between 26  C and 31  C, the short-time
variation band being around a maximum of 3  C.
To determine the influence of temperature on the operating performance of a
desalination plant, for example, on the bandwidth and annual average of its energy
consumption, it is necessary to know the annual distribution of the seawater
90 3 Seawater: Composition and Properties

temperature. Such a temperature distribution profile, i.e. the number of hours per
year and the percentage hours per year for each seawater temperature, is presented in
Fig. 3.8 for a location in the Arabian Gulf. Knowledge of the length of periods of
high temperature can also be useful for predicting the times of possibly increased
biological growth in the seawater and thus the risk of malfunction in the pretreatment
system of a seawater desalination plant.

Fig. 3.7 Seasonal seawater temperature distribution at locations in Arabian Gulf and Singapore

Fig. 3.8 Annual seawater temperature distribution at an Arabian Gulf location


3.2 Properties of Seawater 91

3.2 Properties of Seawater

Designing a seawater desalination plant and calculating the relevant technical


systems require knowledge of a number of thermophysical, physical, and physico-
chemical properties of the seawater and their dependence on salt content, tempera-
ture, and sometimes also pressure (Table 3.4).
According to the design and operating conditions of a desalination plant, the
range of validity of the algorithms used to determine the values of the specified
parameters should be:

• For salinity up to a maximum of 100 g/kg


• For temperature for thermal desalination processes up to a maximum of 70  C for
MED, up to a maximum of 120  C for MSF, and up to a maximum of 40  C for
membrane desalination.

Table 3.4 Relevant seawater properties for seawater desalination design


No. Category of properties Property
1 Thermophysical
1.1 Specific heat capacity
1.2 Specific enthalpy
1.3 Thermal conductivity
1.4 Vapour pressure
1.5 Boiling point elevation
1.6 Latent heat of evaporation
2 Physical
2.1 Density
2.2 Dynamic viscosity
2.3 Surface tension
2.4 Solubility of gases
2.4.1 • Oxygen
2.4.2 • Nitrogen
2.4.3 • Carbon dioxide
3 Physico-chemical
3.1 Ionic strength
3.2 Activity coefficients, solubility of compounds
3.3 Osmotic coefficient and osmotic pressure
3.4 Bicarbonate/carbonate/CO2 equilibrium
3.5 Boric acid/borate/pH equilibrium
92 3 Seawater: Composition and Properties

The operating pressure of a reverse osmosis seawater desalination plant is


normally in the range between 60 and 80 bar, although it can sometimes be up to
110–120 bar. The evaporator of a thermal process is operated at a pressure between
less than 0.1 and a maximum of 2 bar. Under these operating conditions, therefore,
pressure has no influence on the specified parameters of the actual desalination
process. Also in the operating pressure range of a reverse osmosis seawater desali-
nation plant, under technical process design aspects, the influence of the pressure
dependence of the thermophysical and physical properties of the process part
streams on the RO process itself can be neglected. However, this is not the case
with respect to the auxiliary systems of a particular desalination process, where the
physical and physico-chemical properties of the seawater and other side-streams
may very well be highly pressure-dependent, for example, regarding the solubility of
oxygen, nitrogen, air, and carbon dioxide.
The various properties of seawater and their dependence on the salt content,
temperature, and pressure of the seawater have been investigated in a multiplicity of
oceanographic measurements with natural and artificial seawater samples. For a
mathematical definition of the relevant properties of seawater and their dependence
on ambient conditions, the thus obtained series of measured data were used to
develop polynomial equations using regression methods. Such measurements to
determine the density of seawater and its dependence on temperature and chlorinity
were conducted by Knudsen as early as at the beginning of the twentieth century.
With various modifications, the resulting equations were used in oceanography up
until the 1970s. More recent measurements, particularly by Millero [9], including at
higher pressure, then indicated the need to develop an oceanographic standard in
which salinity was determined no longer by chlorinity, but by conductivity measure-
ment. The thus determined salinity, referred to at that time as practical salinity Sp,
formed the new basis for determining the salt content of seawater. Also, the
dependence of seawater properties on seawater salt content was then determined
on the basis of practical salinity and not chlorinity.

Practical Salinity Scale (PSS-78) and International Equation of State


for Seawater (EOS80)
In the period from 1978 to 1981, the Practical Salinity Scale (PSS-78) was prepared
and described in various publications and UNESCO Technical Papers in Marine
Science [3, 4]. Subsequently, in 1983, with UNESCO Technical Papers in Marine
Science 44, this standard was introduced for the determination of salinity by means
of conductivity measurement (see also Sect. 3.1.1) and, based thereon, the Interna-
tional Equation of State for Seawater (EOS80) as a standard for the calculation of
various properties of seawater and their dependence on pressure, salinity, and
temperature [5]. Of the above-mentioned properties of seawater, however, this
oceanographic standard could be used to determine only the density and specific
heat of seawater and even then only in a range of validity of salinity from 0 to 42 g/
kg and of temperature from 2 to 40  C.
3.2 Properties of Seawater 93

International Thermodynamic Equation of Seawater-2010 (TEOS-10)


From 2005 onwards began the development of a comprehensive standard based on
thermodynamic principles for the properties of seawater in its various states of
aggregation (water, ice, seawater, and seawater/humid air). In 2008, this standard
was published as Release IAPWS 08 and Supplementary Release IAPWS 09 of the
International Association for the Properties of Waters and Steam (IAPWS) [10, 11]
and was introduced in 2010 as the International Thermodynamic Equation of
Seawater 2010 (TEOS-10). This standard superseded the EOS80 standard. A signif-
icant part of the new standard was the definition of SR and, based thereon, SA. In
TEOS-10, absolute salinity superseded practical salinity as an input parameter in
algorithms for calculating the properties of seawater (see also Sect. 3.1.1).
TEOS-10 combined the results of practical measurements for determining sea-
water properties with thermodynamic principles and laws for the various states of
aggregation of seawater. The thermodynamic basis of TEOS-10 is a function of the
thermodynamic potential describing the thermodynamic status of seawater in its
various states of aggregation. This so-called Gibbs function can be used to determine
all the thermodynamic properties of seawater directly or from a combination of
derivations of the function. It is an empirical function of absolute salinity, tempera-
ture, and pressure. Thus, density, specific heat capacity, and specific enthalpy can be
calculated directly from the Gibbs function, while other properties of seawater can be
calculated from its derivations.
The contents of the standard and its bases have been described in a multiplicity of
documents [12–15]. Software tools are made available to users in the form of
calculating aids for the highly extensive and complex equation systems and calcula-
tion processes. Documents and software packages can be downloaded from the
standard’s own homepage (www.TEOS-10.org).
The software contained in TEOS-10 consists of two packages:

• The Seawater-Ice-Air (SIA) library


• The Gibbs SeaWater (GSW) Oceanographic Toolbox

The two software packages provide library routines for calculating the multiplic-
ity of parameters, properties, and constants contained in TEOS-10 as well as for
converting units of measure for pressure, temperature, and salinity.
The SIA library contains the entire scope of calculations both for water and
seawater in their various states of aggregation in the form of liquid and ice and
also for the corresponding equilibria between water, ice, and air. The library is
described and documented in detail in [14, 15].
The GSW Toolbox is limited to the properties and calculation routines for
seawater in liquid phase. A detailed description of the Toolbox is available in
[13]. It is set up for use in oceanography and differs also with regard to input from
the routines in the SIA library, where the required parameters for SA, temperature,
and pressure are entered as follows:
94 3 Seawater: Composition and Properties

• Absolute salinity (SA) in [kg/kg]


• Temperature (T ) in [K]
• Pressure (P) in Pascal [Pa] and as absolute pressure

The GSW Toolbox adapts these parameters to units that are more customary in
oceanography, entered as follows:

• Absolute salinity (SA) in [g/kg]


• Temperature (t) in [ C]
• Pressure ( p) in [dbar] and as standard pressure

Temperature (T) in K is calculated from t in  C as follows Eq. (3.26):

T ¼ t þ 273:15 ð3:26Þ

T ¼ absolute temperature [K]


t ¼ Celsius temperature [ C]

and standard pressure is calculated as follows:

p ¼ P  P0 ð3:27Þ

p ¼ atmospheric pressure [Pa]


P ¼ absolute pressure [Pa]
P0 ¼ absolute pressure of one standard atmosphere ¼ 101,325 [Pa]

With regard to conversion from Pascal Pa to decibar (dbar), reference is made to


Annex 3.A3, Table 3.35, in the appendix to this chapter.

Table 3.5 Properties of seawater routines in SIA software package of TEOS-10


Property SIA module SIA function
Specific heat capacity Sea_3a sea_cp_si
GSW_Library_5 gsw_cp
Specific enthalpy Sea_3a sea_enthalpy_si
GSW_Library_5 gsw_enthalpy
Vapour pressure Sea_Vap_4 sea_vap_vapourpressure_si
Specific heat of evaporation Sea_Vap_4 sea_vap_enthalphy_evap_si
Density Sea_3a sea_density_si
GSW_Library_5 gsw_dens
Osmotic coefficient Sea_3a sea_osm_coeff_si
Osmotic pressure Sea_Liq_4 sea_liq_osmoticpressure_si
3.2 Properties of Seawater 95

The GSW Toolbox makes the library routines available in the Fortran and
C programming languages and additionally as routines for MATLAB. The routines
in the SIA package are programmed in Fortran and Visual Basic. The routines can be
integrated into a suitable program environment, after which they make the
corresponding calculation processes available in the form of functions. With the
Visual Basic software package, it is possible also to import the individual modules
into the Excel spreadsheet program. The calculation programs are then available as
functions in the worksheets. The SIA software package additionally contains parts of
the GSW Toolbox with the same method of input in units as used in the Toolbox. As
far as integration into a corresponding program environment is concerned, it must be
ensured that the additional modules and functions required for execution of the
relevant function are available. The modules and subfunctions belonging to each
function are described in detail in the documentation of the two software packages.
TEOS-10 contains part of the thermophysical parameters given in Table 3.4
(No. 1.1–1.3, 1.4, and 1.6) as well as the physical parameter of density (No. 2.1)
and the physico-chemical properties of osmotic coefficient and osmotic pressure
(No. 3.3). The range of validity of the equations at atmospheric pressure extends
from 0 to 120 g/kg (0 to 0.12 kg/kg) for salinity and in a range 6–80  C and
263.15–353.15 K, for temperature, respectively. To calculate the density, it is
possible to extend the temperature range up to 90  C and salinity up to a maximum
of 70 g/kg.
Table 3.5 lists the function names and software modules under which some of the
above-mentioned properties of seawater can be found in the SIA library and its GSW
5 library module.
The technical calculation of seawater desalination processes and their relevant
systems requires knowledge of the properties of the seawater in the liquid state and
their dependence on salinity and temperature in the above-specified ranges (salinity
up to 100 g/kg, temperature up to 70  C/120  C for thermal processes and up to
40  C for reverse osmosis). At the predominant operating pressures in seawater
desalination and given the required accuracy for technical calculations, the pressure
dependence of the thermophysical and physical properties (with the exception of gas
solubility) can be left out of consideration. Also the complex thermodynamic
equation systems in TEOS-10 for fully describing the various states of aggregation
of seawater are not absolutely necessary for the technical design of a seawater
desalination plant. When it is additionally considered that the TEOS-10 equations
do not fully cover the range of parameters necessary for SWRO design, it is suitable,
when engineering such processes, to apply the calculation algorithms derived
directly from measurements and investigations with seawater as described below
(see Sects. 3.2.1, 3.2.2, and 3.2.3).
96 3 Seawater: Composition and Properties

A first comprehensive description of the thermophysical, physical, and physico-


chemical properties of seawater for the purpose of the technical calculation of
seawater desalination plants was provided by H. E. Hömig in 1978 [16]. This
work, a handbook for the design of thermal seawater desalination plants, especially
multi-stage flash distillation (MSF) plants, covered most of the seawater properties
given in Table 3.4, with the exceptions of osmotic pressure, osmotic coefficient
(No. 3.3), and boric acid/borate equilibrium (No. 3.5), parameters of importance for
the design of a reverse osmosis plant. The therein described algorithms are based on
investigations and research results up to 1975. The range of validity of the equations
for calculating the parameters is from 0 to 80 g/kg for salinity and from 20 to 120  C
for temperature. Nitrogen solubility is given for a chlorinity range from 15 to 21 g/kg
and a temperature range from 0 to 28  C, with oxygen solubility being given for a
chlorinity range from 0 to 30 g/kg and a temperature range from 5 to 35  C.
A further comprehensive and detailed presentation of the level of knowledge with
regard to the thermophysical and physical properties of seawater as required for the
design of seawater desalination plants, as well as their calculation and dependence
on salinity and temperature was published by Sharqawy et al. in 2010 [17]. This
work includes a presentation and comparison of the calculation algorithms published
in the period from 1920 to 2008 on the basis of seawater measurements. It also
contains an assessment of the extent to which the results of the calculation equations
are consistent with the data obtained with Release IAPWS 08 and therefore also with
TEOS-10. The work additionally examines whether the calculation results at zero
salinity agree with the properties of normal water according to the IAPWS-95
standard [18, 19]. The publication covers the thermophysical and physical properties
given as No. 1.1 to 2.3 in Table 3.4 as well as calculation of the osmotic coefficient
(No. 3.3). A calculation equation is recommended for each of the parameters. The
range of validity of the algorithms is in all cases in the salinity range 0–120 g/kg or
0–0.12 kg/kg and in the temperature range 0–120  C.
The following compilation of calculation equations and values relies mainly on
[16, 17] for the thermophysical and physical properties of seawater (No. 1.1–2.4 in
Table 3.4) and mainly on [16] and more recent publications for the physico-chemical
properties (No. 3.1–3.5 in Table 3.4).
The following gives the calculation algorithm for each of the individual property
parameters. In addition, the dependence of the parameter on temperature and salinity
is presented in a graph. Respective data tables can be found in Annex 3.A2,
Tables 3.17–3.26 of this chapter. Also given for the individual algorithms is their
range of validity in the respective input units.
The heading for each individual parameter indicates the main reference source.
References to the original sources of the calculation equations, not the main source,
are contained in the footnotes.
3.2 Properties of Seawater 97

3.2.1 Thermophysical Properties

3.2.1.1 Specific Heat Capacity [16]


 
cp,SW ¼ 0:001  A þ B  t þ C  t 2 þ D  t 3 ð3:28Þ

A ¼ 4206:8  6:6197  S þ 1:2288  102  S2

B ¼ 1:1262 þ 5:4178  102  S  2:2719  104  S2

C ¼ 1:2026  102  5:3566  104  S þ 1:8906  106  S2

D ¼ 6:8774  107 þ 1:5170  106  S  4:4268  109  S2

Range of validity of Eq. (3.28): S ¼ 0–160 g/kg; t ¼ 0–180  C


Cp,SW ¼ specific heat capacity [kJ/kg,K], see (Fig. 3.9)
t ¼ temperature [ C]
S ¼ salinity (SA, SR, TDSM) [g/kg]

Fig. 3.9 Specific heat capacity of seawater


98 3 Seawater: Composition and Properties

3.2.1.2 Specific Enthalpy [17]



hsw ¼ hw  0:001  S  a1 þ a2  S þ a3  S2 þ a4  S3 þ a5  t þ a6

t 2 þ a7  t 3 þ a8  S  t þ a9  S2 þ a10  S  t 2
ð3:29Þ

hW ¼ 141:355 þ 4, 202:07  t  0:535  t 2 þ 0:004  t 3 ð3:30Þ

a1¼2.348  104 a6¼+4.417  101


a2¼+3.152  105 a7¼+2.139  101
a3¼+2.803  106 a8¼1.991  104
a4¼1.446  107 a9¼+2.778  104
a5¼+7.826  103 a10¼+9.728  101

Range of validity of Eq. (3.29): S ¼ 0–0.12 kg/kg; t ¼ 10–120  C


Range of validity of Eq. (3.30): t ¼ 5–200  C
hSW ¼ specific enthalpy of seawater [kJ/kg], see (Fig. 3.10)
hW ¼ specific enthalpy of pure water [kJ/kg]
S ¼ salinity (SA, SR, TDSM) [kg/kg]
t ¼ temperature [ C]

Fig. 3.10 Specific enthalpy of seawater


3.2 Properties of Seawater 99

Fig. 3.11 Thermal conductivity of seawater

3.2.1.3 Thermal Conductivity [17]


log 10 ðk SW Þ ¼ log 10 ð240 þ 0:0002  SÞ þ 0:434
 
343:5 þ 0:037  S
 2:3 
t þ 273:15
 0:333
t þ 273:15
 1 ð3:31Þ
647 þ 0:03  S

Range of validity of Eq. (3.31): S ¼ 0–160 g/kg; t ¼ 0–180  C1


kSW ¼ thermal conductivity of seawater [mW/mK], see (Fig. 3.11)
S ¼ salinity (SA, SR, TDSM) [g/kg]
t ¼ temperature [ C]

3.2.1.4 Vapour Pressure [16, 17]


   
log 10 pV,SW ¼ log 10 pV,W  2:1609  104  S  3:5012  107
 S2 ð3:32Þ

1
D.T. Jamieson and J.S. Tudhope, Physical properties of seawater solutions—Thermal Conductiv-
ity, Desalination, 8 (1970), pp. 393–401.
100 3 Seawater: Composition and Properties

pV,SW ¼ 105  10 log 10 ðpV,SW Þ ð3:33Þ

Range of validity of Eq. (3.32): S ¼ 0–160 g/kg2


 
log 10 pV,W ¼ 0:434295
 
a
 1 þ a2 þ a3  T þ a4  T 2 þ a5  T 3 þ a6  ln ðT Þ
T
ð3:34Þ

a1 ¼ 5800221  103 a4 ¼ +4.17648  105


a2 ¼ +1.391499 a5 ¼ 1.44521  108
a3 ¼ 4.864024  102 a6 ¼ +6.54597

Range of validity of Eq. (3.34): T ¼ 273.15–473.15 K ¼ 0–200  C3


Rough calculations can be performed using the simpler Eq. (3.35), which yields
good approximate values for pv,sw, especially for standard seawater:

pV,SW ¼ pV,W  ð1  0, 000537  SÞ ð3:35Þ

PV,SW ¼ vapour pressure of seawater [bar], see (Fig. 3.12)


PV,W ¼ vapour pressure of pure water [bar]
T ¼ temperature [K]
S ¼ salinity [g/kg]

3.2.1.5 Boiling Point Elevation [17]


BPE ¼ A  S2 þ B  S ð3:36Þ

A ¼ 4:58385  104  t 2 þ 2:823095  101  t þ 17:9452

B ¼ 1:536175  104  t 2 þ 5:26691  102  t þ 6:5605

BPE ¼ boiling point elevation [K], see (Fig. 3.13)


t ¼ temperature [ C]
S ¼ salinity [kg/kg]

2
W.H. Emerson and D. Jamieson, Some physical properties of seawater in different concentrations,
Desalination, 3 (1967), pp. 207–212.
3
ASHRAE Handbook: Fundamentals, American Society of Heating, Refrigerating,
Air-Conditioning Engineers, and Inc., ASHRAE, 2005.
3.2 Properties of Seawater 101

Fig. 3.12 Vapour pressure of seawater

Fig. 3.13 Boiling point elevation of seawater


102 3 Seawater: Composition and Properties

Fig. 3.14 Latent heat of vaporization of seawater

3.2.1.6 Latent Heat of Vaporization [17]


 
S
hFG,SW ¼ hFG,W  1  ð3:37Þ
1000
 
hFG,W ¼ 0:001  a1 þ a2  t þ a3  t 2 þ a4  t 3 þ a5  t 4 ð3:38Þ

a1 ¼ +2.5009  106 a4 ¼ 8.1028  103


a2 ¼ 2.3692  103 a5 ¼ 2.0799  105
a3 ¼ +2.6776  101

Range of validity of Eq. (3.38): t ¼ 0–200  C


hFG,SW ¼ latent heat of vaporization of seawater [kJ/kg], see (Fig. 3.14)
hFG,W ¼ latent heat of vaporization of pure water [kJ/kg]
t ¼ temperature [ C]
S ¼ salinity [g/kg]
3.2 Properties of Seawater 103

3.2.2 Physical Properties

3.2.2.1 Density [17]


ρSW¼ ρW
 
þ b1  S þ b2  S  t þ b3  S  t 2 þ b4  S  t 3 þ b 5  S2  t 2 ð3:39Þ

ρW ¼ a1 þ a2  t þ a3  t 2 þ a4  t 3 þ a5  t 4 ð3:40Þ

a1 ¼ +9.9992  102 b1 ¼ +8.0200  102


a2 ¼ +2.0341  102 b2 ¼ 2.0005
a3 ¼ 6.1625  103 b3 ¼ 1.6771  102
a4 ¼ +2.2615  105 b4 ¼ 3.0601  105
a5 ¼ 4.6571  108 b5 ¼ 1.6132  105

Range of validity of Eq. (3.39): S ¼ 0–0.16 kg/kg ¼ 0–160 g/kg; t ¼ 0–180  C

Range of validity of Eq. (3.40): t ¼ 0–180  C


ρSW ¼ density of seawater [kg/m3], see (Fig. 3.15)
ρW ¼ density of pure water [kg/m3]
S ¼ salinity [kg/kg]
t ¼ temperature [ C]

Fig. 3.15 Density of seawater


104 3 Seawater: Composition and Properties

3.2.2.2 Dynamic Viscosity [17]


 
μSW ¼ μW  1 þ A  S þ B  S2 ð3:41Þ

A ¼ 1:541 þ 1:998  102  t  9:52  105  t 2

B ¼ 7:974  7:5615  102  t þ 4:7237  104  t 2


 1
μW ¼ 4:2844  105 þ 0:157  ðt þ 64:993Þ2  91:2965 ð3:42Þ

Range of validity of Eq. (3.41): S ¼ 0–0.15 kg/kg ¼ 0–150 g/kg; t ¼ 0–180  C


Range of validity of Eq. (3.42): t ¼ 0–180  C
μSW ¼ dynamic viscosity of seawater [kg/m.s], see (Fig. 3.16)
μW ¼ dynamic viscosity of pure water [kg/m.s]
S ¼ salinity [kg/kg]
t ¼ temperature [ C]

3.2.2.3 Kinematic Viscosity


Kinematic viscosity is calculated from dynamic viscosity and density as follows:

Fig. 3.16 Dynamic viscosity of seawater


3.2 Properties of Seawater 105

Fig. 3.17 Kinematic viscosity of seawater

μSW
νSW ¼ ð3:43Þ
ρSW

νSW ¼ kinematic viscosity of seawater [m2/s], see (Fig. 3.17)


ρSW ¼ density of seawater [kg/m3] (Fig. 3.15)

3.2.2.4 Surface Tension [17]


σ SW ¼ σ W  ð1 þ ð0:0002264  t þ 0:009458Þ  ln ð1 þ 0:03311  SÞÞ ð3:44Þ
 1:256
t þ 273:15
σW ¼ 0:2358  1 
647:096
 
t þ 273:15
 1  0:625  1  ð3:45Þ
647:096

Range of validity of Eq. (3.44): S ¼ 0–40 g/kg; t ¼ 0–40  C4


Range of validity of Eq. (3.45): t ¼ 0–370  C

4
International Association for the Properties of Water and Steam, Release on the IAPWS Surface
Tension of Ordinary Water Substance, 1994.
106 3 Seawater: Composition and Properties

Fig. 3.18 Surface tension of seawater

σ SW ¼ surface tension of seawater [N/m], see (Fig. 3.18)


σ W ¼ surface tension of pure water [N/m]
S ¼ salinity [g/kg]
t ¼ temperature [ C]

3.2.2.5 Solubility of Gases in Seawater


All the gases present in the atmosphere are contained in dissolved form in varying
quantities in seawater, where, according to Henry’s law, the quantity of dissolved
gas (cg) is proportional to its partial pressure (Pg) Eq. (3.46):

P g ¼ H g  cg ð3:46Þ

Pg ¼ pressure of gas [atm]


cg, ¼ concentration of gas in solvent (water, seawater) [mol/l]
Hg ¼ Henry’s coefficient [atm.l/mol]

The solubility of a given gas is characterized by Henry’s coefficient (Hg)


Eq. (3.47):

Pg
cg ¼ ð3:47Þ
Hg
3.2 Properties of Seawater 107

For a given gas with freshwater as solvent, the value of Henry’s coefficient is
dependent exclusively on temperature. With seawater as the solvent, however, the
value is additionally dependent on salinity.
The dimension of Hg is determined by the choice of units for the concentration of
the gas in solution and in the gaseous phase. For the purposes of technical
calculations, it is usual to give the concentration of the gas in solution in mol/l and
in the gaseous phase as partial pressure in atm or bar, in which case Henry’s
coefficient has the dimension atm  l/mol or bar  l/mol. In technical calculations,
however, it is easier to use the reciprocal value of Henry’s coefficient as solubility
coefficient (hg,v), as this then has the dimension of a gas concentration mol/l  atm
(bar) Eq. (3.48). Henry’s equation Eq. (3.47) then becomes Eq. (3.49):

1
hg ¼ ð3:48Þ
Hg

c g ¼ hg  P g ð3:49Þ

The solubility of gases is frequently described also by the Bunsen absorption


coefficient (βg). βg is that volume of gas in litre at a temperature of 0  C and a
pressure of 1 atm that is dissolved in 1 l of solvent at a partial pressure of the gas of
1 atm. βg therefore has the dimension l/l, atm. The solubility coefficient (hg,v) is
calculated from the Bunsen coefficient Eq. (3.50):

βg
hg,V ¼ ð3:50Þ
V M,g

hg,v ¼ solubility coefficient by volume [mol/l.atm]


βg ¼ Bunsen absorption coefficient [l/l.atm]
VM,g ¼ molar volume of gas [l/mol] ¼ ~22.398 for O2; ~22.41 for N2

Oxygen and Nitrogen


The value of the Bunsen absorption coefficient of oxygen βO2 or nitrogen βN2 for
seawater of a given salinity and temperature can be calculated with Eq. (3.51) [20]:
   
100 t þ 273:15
ln βO2,N ¼ A1 þ A2  þ A3  ln þS
2 t þ 273:15 100
"    2 #
t þ 273:15 t þ 273:15
 B1 þ B2  þ B3  ð3:51Þ
100 100

Gas A1 A2 A3 B1 B2 B3
Oxygen 58.3877 +85.8079 +23.8439 0.034892 +0.015568 0.0019387
(O2)
Nitrogen 59.6274 +85.7661 +24.3696 0.051580 +0.026329 0.0037252
(N2)
108 3 Seawater: Composition and Properties

Fig. 3.19 Oxygen solubility coefficient in seawater

With Eq. (3.50), this can then be used to calculate the solubility coefficient hO2 ,v
for oxygen and hN2 ,v for nitrogen. The dependence of the solubility coefficients of
both gases on salinity and temperature is presented in graphs in Figs. 3.19 and 3.20
for a salinity range of 0–40 g/kg and a temperature range of 10–40  C.
The associated data are contained in Annex 3.A2, Tables 3.27 and 3.28. The
given solubility coefficients are valid for dry gas.
To determine the concentration of the gas in the liquid phase in the unit of
concentration mg/l, as is usual for technical applications, the solubility coefficient
of the gas hg,v is multiplied as in Eq. (3.52) by its molecular weight (Mg) and pressure
(Pg) (see also Eq. 3.53 for oxygen and Eq. 3.54 for nitrogen):

cg ¼ hg,v  Pg  MWg  103 ½mg=l ð3:52Þ

cO2 ¼ 32  103  hO2 ,V  PO2 ½mg=l ð3:53Þ

cN2 ¼ 28:01  103  hN2 ,V  PN2 ½mg=l ð3:54Þ

MWg ¼ molecular weight of gas ¼ 32.0 for O2, 28.01 for N2


3.2 Properties of Seawater 109

Fig. 3.20 Nitrogen solubility coefficient in seawater

For a gas pressure range of oxygen and nitrogen of 1–20 bar and a seawater
temperature of 20  C, Fig. 3.21 shows the gas solubility in mg/l for a seawater
salinity range of 0–40 g/kg as calculated according to the above-described method.
In a gas mixture, the gas solubility is calculated not with the pressure of the pure
gas Pg but with its partial pressure Pg,i in the gas mixture.
The oxygen saturation value of seawater in contact with humid atmospheric air at
USAC (unit standard atmospheric concentrations) conditions, i.e. at a partial pres-
sure of oxygen (PO2 ) of 0.2095 atm and of nitrogen (PN2 ) of 0.781 atm, is, therefore,
as shown above for the solubility coefficients of the gases, dependent not only on the
temperature of the seawater but also on its salinity. The oxygen saturation concen-
tration (C O2,atm ) in mg/l for a given salinity and temperature of seawater in contact
with the atmosphere is calculated as in Eq. (3.55) [21]. The effects of these
conditions on the oxygen saturation of seawater are presented in a graph in Fig. 3.22.

a2 a3 a4
ln cO2,atm ¼ a1 þ þ þ
ð273:15 þ t Þ ð273:15 þ t Þ2 ð273:15 þ t Þ3
!
a5 b2 b3
þ S b1 þ þ ð3:55Þ
ð273:15 þ t Þ4 ð273:15 þ t Þ ð273:15 þ t Þ2
110 3 Seawater: Composition and Properties

Fig. 3.21 Oxygen and nitrogen concentration in seawater up to 20 bar at 20  C

a1 ¼ 1.3934411  102 a5 ¼ 8.621949  1011


a2 ¼ +1.575701 105 b1 ¼ +17674  102
a3 ¼ 6.642308  107 b2 ¼ 10.754
a4 ¼ +1.243800  1010 b3 ¼ +2.1407103

CO2,atm ¼ oxygen concentration [mg/l] at USAC (unit standard atmospheric


concentrations)

Strictly speaking, the proportionality of Henry’s law applies only to a pressure


range in which gases still behave ideally. If this range is exceeded, then, when
determining the value of the Henry coefficient or solubility coefficient of a gas, it is
necessary to take into consideration not only the influence of the temperature and salt
content (salinity) of the solvent but also the pressure of the gas. The value of the
coefficient changes with increasing pressure. Technical calculations should take into
consideration the influence of pressure on the solubility coefficient of oxygen from
around 80 to 100 bar. For nitrogen, this range begins already at around 50 bar.
To determine the solubility of either of the two gases or of both together (when
determining the solubility of air) for seawater or concentrates at gas pressures in
excess of the above-mentioned range, it is necessary to employ more complex
3.2 Properties of Seawater 111

calculation algorithms. Although such polynomial equations also contain calculation


parameters for temperature and pressure, the salinity of seawater is no longer the sole
calculation parameter with regard to the influence of salt content on gas solubility, it
being additionally necessary to take into consideration the concentration of each
individual component of the salt content [22, 23]. The range of validity of these
algorithms extends the temperature range of 0–40  C in Eq. (3.55) to above 300  C.
However, in the pressure and temperature ranges in which membrane seawater
desalination plants are designed and operated, it is not normally necessary to take
into consideration the non-ideal behaviour of oxygen and nitrogen.

Carbon Dioxide
Equation (3.56) can be used to determine the solubility coefficient hCO2 ,V of carbon
dioxide as a function of salinity and temperature [24]:
   
100 273:15 þ t
ln hCO2 ,V,M ¼ A1 þ A2  þ A3  ln þS
273:15 þ t 100
"    2 #
273:15 þ t 273:15 þ t
 B1 þ B2  þ B3  ð3:56Þ
100 100

Fig. 3.22 Oxygen saturation concentration of seawater at USAC conditions


112 3 Seawater: Composition and Properties

Unit of
hCO2 ,V,M A1 A2 A3 B1 B2 B3
mol/l.atm 58.0931 +90.5069 +22.2940 0.027766 0.025888 +0.0050578
mol/kg.atm 60.2409 +93.4517 +23.3585 0.023517 0.023656 +0.0047036

hCO2 ,V ¼ solubility coefficient of carbon dioxide by volume [mol/l.atm]


hCO2 ,M ¼ solubility coefficient of carbon dioxide by mass [mol/kg.atm]

The solubility coefficients hCO2 ,V of CO2 calculated for the salinity range of
0–40 g/kg and temperature range of 0–40  C are presented in Fig. 3.23 for those
parameter ranges. The associated data are given in Annex 3.A2, Table 3.29.
The carbon dioxide concentration of CO2 in seawater is calculated from its
solubility coefficient as in Eq. (3.57), as already described above for oxygen and
nitrogen:

cCO2 ¼ 44, 01  103  hCO2 ,V  PCO2 ½mg=l ð3:57Þ

Figure 3.24 presents the carbon dioxide concentration for the pressure range of
the gas from 1 to 20 bar at a constant seawater temperature of 20  C, albeit with
varying salinity in the range of 0–40 g/kg.

Fig. 3.23 Carbon dioxide solubility coefficient of seawater


3.2 Properties of Seawater 113

Fig. 3.24 Carbon dioxide concentration in seawater up to 20 bar at 20  C

For carbon dioxide, technical calculations should take into consideration the
influence of pressure on the solubility coefficient starting from a pressure range of
20–30 bar. As already explained above for nitrogen and oxygen, it is then necessary
to employ more complex calculation algorithms than those in Eq. (3.56), which, to
determine the gas solubility, take into consideration not only pressure and tempera-
ture but also the more specific composition of the salt content of the seawater or
concentrate [25]. Also, the temperature range of 0–40  C in Eq. (3.56) for calculating
the CO2 solubility coefficient is extended to above 250  C. Yet also in this case, the
use of Eq. (3.56) is normally sufficient for reverse osmosis seawater desalination and
its usual operating conditions.

3.2.3 Physico-Chemical Properties

Like the physical and thermophysical parameters, the physico-chemical properties of


seawater are also influenced by its salt content. In this case, however, salinity cannot
be used as the sole parameter. To calculate the physico-chemical properties, it is
necessary to take into consideration in greater detail the specific composition of the
seawater salt content and its main components. The primary unit of concentration of
the salt content components is their molar concentration in the form of molarity per
mass (mMi,m) or molarity per volume (mMi,v) of solution, i.e. seawater (molarity), or
molarity per mass of solvent (mmi), i.e. pure water H2O (molality). These molar
114 3 Seawater: Composition and Properties

concentrations are calculated from the concentrations in units of mass and volume as
follows:

ci,v
mMi,v ¼ ð3:58Þ
MWi
ci,m
mMi,m ¼ ð3:59Þ
MWi
ci,m,H2 O
mmi ¼ ð3:60Þ
MWi

mMi,v ¼ molarity per volume ¼ mol per litre of solution [mol/l]


mMi,m ¼ molarity per mass ¼ mol per kg of solution [mol/kg]
mmi ¼ molality ¼ mol per kg of solvent (water, H2O) [mol=kgH2 O ]
MWi ¼ molecular weight of component i [g/mol]
ci,v ¼ solution concentration of component i per volume [g/l]
ci,m ¼ solution concentration of component i per mass [g/kg]
ci,m,H2 O ¼ mass per solvent (H2O) concentration of component i [g=kgH2 O ]

ci,v 1000
mmi ¼  ð3:61Þ
ϱSW  MWi 1000  S

S ¼ SR ¼ SA ¼ salinity [g/kg] ¼ solution-based salinity

The molality of a component i is calculated from the concentration unit ci in


[mg/l], which is the normal unit in desalination technology, as follows:

ci 106
mmi ¼  6 ð3:62Þ
1000  ϱSW  MWi 10  TDSM

ci ¼ concentration of component i per volume [mg/l]


TDSM ¼ total dissolved solids content, mass-based [mg/kg]

Additional algorithms for the calculation of molar concentrations from mass


concentrations and vice versa are contained in Annex 3.A3, Table 3.36.
The molar concentration for reference seawater is 1.12 mol/kg based on seawater
and 1.16 mol=kgH2 O based on solvent, i.e. H2O (see also Table 3.6).
Table 3.6 Molar concentration and ionic strength of reference seawater
Valence of ions Atomic/molecular weight Concentration
Zi AWi/MWi ci,v mMi,m Zi2mMi,m mmi Zi2mmi
Component Symbol _ g/mol g/kgSW g-mol/kgSW g-mol/kgSW g‐mol=kgH2 O g‐mol=kgH2 O
3.2 Properties of Seawater

Calcium Ca++ 2 40.078 0.41208 0.01028 0.04113 0.010657 0.04263


Magnesium Mg++ 2 24.305 1.28372 0.052817 0.21127 0.054742 0.21897
Sodium Na+ 1 22.990 10.78145 0.468967 0.46897 0.486059 0.48606
Potassium K+ 1 39.098 0.39910 0.010208 0.01021 0.010580 0.01058
Strontium Sr++ 2 87.620 0.00795 0.000091 0.00036 0.000094 0.00038
Carbonate CO3 2 60.009 0.01434 0.000239 0.00096 0.000248 0.00099
Bicarbonate HCO3 1 61.017 0.10481 0.001718 0.00172 0.001780 0.00178
Sulphate SO4 2 96.063 2.71235 0.028235 0.11294 0.029264 0.11706
Chloride Cl 1 35.453 19.35271 0.545869 0.54587 0.565764 0.56576
Bromide Br 1 79.904 0.06728 0.000842 0.00084 0.000873 0.00087
Boron B(OH)3 0 61.833 0.01944 0.000314 0.00000 0.000326 0.00000
[B(OH)4] 1 78.840 0.00795 0.000101 0.00010 0.000105 0.00010
Fluoride F 1 18.998 0.0013 0.000068 0.00007 0.000071 0.00007
Hydroxyl OH 1 17.007 0.00014 0.000008 0.00001 0.000009 0.00001
Total ∑ 35.165 1.120 1.394 1.161 1.445
Ionic strength Im – – – – – – 0.723
IM,m – – – – 0.697 – –
115
116 3 Seawater: Composition and Properties

3.2.3.1 Ionic Strength


The ionic strength Im or IM,m of seawater is calculated from the molar concentrations
of the components of the seawater salt content in the form of molarity or molality as
in Eqs. (3.63) or (3.64), respectively. The molarity or molality is summated over all
components of the seawater salt content (see Table 3.6). Any uncharged components
contained in the solution are left out of consideration in the summation.

1 X
i¼n
Im ¼  m  Z 2i ð3:63Þ
2 i¼1 mi

1 X
i¼n
I M,m ¼  m  Z 2i ð3:64Þ
2 i¼1 M,mi

Im ¼ ionic strength, solvent-based [mol=kgH2 O ]


IM,m ¼ ionic strength, solution-based [mol/kg solution]
Zi ¼ charge (valence) of component
i ¼ individual component identification number

To calculate the physico-chemical properties of seawater and also chemical


reactions and equilibria in the sea and their influencing by seawater salt content,
salinity is mainly replaced by ionic strength (Im), which is calculated from molality
(mmi).
For reference seawater, this yields an Im of 0:723mol=kgH2 O and IM,m of
0.697 mol/kg of seawater. For seawater of a composition conforming to the principle
of constant proportions and assuming that the main seawater components are fully
dissociated in the form of ions, Fig. 3.25 presents the dependence of the two types of
ionic strength on the salinity of the seawater.
Using the therefrom derived relation as in Eq. (3.65), it is possible, for such
seawater, to calculate good approximations for molarity-based IM,m from seawater
salinity (S):

I M,m ¼ 1:9827  102  S ð3:65Þ

Molality-based Im is then calculated:

1000
I m ¼ I M,m  ð3:66Þ
1000  S
However, such simplified relations should be used only to determine guideline
values for the ionic strength of seawater conforming to the rule of constant
proportions. Concentrates, such as those from a reverse osmosis desalination plant,
are subject to an enrichment of divalent ions, which, according to the algorithm for
calculating the ionic strength, considerably increase the value of the ionic strength
compared with standard seawater owing to potentiation of the ionic charge (see
3.2 Properties of Seawater 117

Table 3.6). The same applies to coastal waters, where water inflow from the land
leads to higher levels of calcium and magnesium than in standard seawater.
Technical calculations often equate Im with molarity-based ionic strength IM,m. As
shown in Fig. 3.25, however, the two types of ionic strength differ significantly from
one another even at medium salinity values and increasingly so at high salinity values.

Ion Association and Ion Pairs


In contrast to freshwater with its lower salt content, the inorganic salt content
components of seawater are not present entirely in the form of ions. The cationic,
positively charged components (CZ+) of the salt content, together with the anionic,
negatively charged components (AZ), can form agglomerates (CAZ+ + Z), of which
the so-called ion pairs are the most important:
þ
þZ 
C Zþ þ nAZ ⇄CAZn

These pairs of ions arise principally from the main components calcium (Ca2+),
magnesium (Mg2+), potassium (K+), sulphate (SO42), carbonate (CO32), and
bicarbonate (HCO3) of the salt content, with sodium (Na+) and chloride (Cl)
taking part only to a lesser extent in ion pair formation.
This results in compounds, some of which are uncharged (neutral) or have a
charge smaller than that of the ions from which they were formed.

Fig. 3.25 Ionic strength of seawater of constant proportions versus salinity


118 3 Seawater: Composition and Properties

The total concentrations of cationic and anionic components in seawater are


composed of the respective concentrations of free ions of the individual components
as well as of the concentrations of the components in the ion pairs (see Eqs. 3.67 and
3.68):
X
½C T  ¼ ½CF  þ ½CA ¼ ½C F  þ ½CAi  ð3:67Þ
i
" #
X
½AT  ¼ ½AF  þ ½ACA  ¼ AF  þ ½AC i  ð3:68Þ
i

[CT] ¼ total molar concentration of cationic component [mol/kg, mol=kgH2 O ]


[AT] ¼ total molar concentration of anionic component [mol/kg, mol=kgH2 O ]
[CF] ¼ molar concentration of free cationic component [mol/kg, mol=kgH2 O ]
[AF] ¼ molar concentration of free anionic component [mol/kg, mol=kgH2 O ]
[CA] ¼ molar concentration of ion pair of components C and A [mol/kg, mol=kgH2 O]
[AC] ¼ molar concentration of ion pair of components A and C [mol/kg, mol=kgH2 O]

For example, the concentration of free calcium ions is calculated from the total
concentration of calcium and the various calcium ion pairs as follows Eq. (3.69):
h i
Caþþ
F ¼ ½CaT   CaCl02  ½CaClþ   CaðSO4 Þ0  CaðHCO3 Þþ

 CaCO03 ð3:69Þ

The extent of ion pair formation is different for the individual components. It is
dependent on the concentration of free cations and anions of the respective partner in
the ion pair and on the concentration of the ion pair itself and is characterized
according to Eq. (3.70) by an equilibrium constant, the stoichiometric association
constant (K CA ).
h þ
i
þZ 
CAZn
n ¼ K CA ð3:70Þ
C Zþ
F  AZ
F

K CA ¼ stoichiometric association constant of ion pair [(mol/kg)1,


(mol/kgH2O)1]
To determine the concentration distribution of the individual salt content
components of seawater into free ions and charged and uncharged ion pairs, it
is necessary to know this association constant for each possible ion-pairing
combination of ions. The various concentration fractions can then be calculated
as follows:
3.2 Properties of Seawater 119

!1
½C F  X n
¼ f CF ¼ 1þ K CAi  AZ
Fi ð3:71Þ
½C T  i

!1
½AF  X
¼ f AF ¼ 1þ K Ci A  CZþ
Fi ð3:72Þ
½AT  i

n
½CA K CAi  AZ
Fi
¼ P n ð3:73Þ
½CT  1 þ K CA  AZ Fi
i
i

½CA K Ci A  C Zþ
Fi
¼ P ð3:74Þ
½AT  1 þ K Ci A  C ZþFi
i

fCF ¼ free cationic ions fraction factor


fAF ¼ free anionic ions fraction factor

The higher the value of the association constant, the higher the fraction of ion
pairs. The constant is temperature-dependent and is additionally influenced by the
salt content, i.e. the ionic strength of the solution [26].
Table 3.7 presents the possible types of ion pairs as originally published by
Garrels and Thompson [27] in connection with their ion-pair model in 1962 and
on the basis of subsequent research results [26, 28–33]. For standard seawater at
25  C, the table additionally presents the percentage band of free ions of the
individual components as well as their fractions in the respective ion pairs according
to various publications.
Especially for sodium and chloride, the difference between the published mini-
mum and maximum values for free ions and ion pairs is considerable. It is nowadays
assumed that, for both of these components, those values tending to the maximum of
free ions are the more accurate, i.e. that sodium in seawater is up to 97–99% in the
form of a free ion and that chloride participates only to a minor extent or almost not
at all in ion pairing, i.e. it is 98–100% in the form of a free ion. The situation is
different with regard to the divalent ions calcium and magnesium as well as, in
particular sulphate and carbonate. The strong oversaturation of seawater
(in comparison with freshwater) with calcium sulphate and calcium carbonate as
well as the increased solubility of alkaline earth sulphates SrSO4 and BaSO4 can be
conclusively explained as a consequence of the formation of neutral ion pairs
through ion association.
In ion association, the charged ions give rise to neutral ion pairs as well as to ion
pairs with a lower charge than the ions that actually participate in pairing. This also
has an influence on the calculation of ionic strength, in which the molality of the
individual solution components is multiplied by the square of their charge. The ionic
strength of a solution decreases according to the extent and nature of ion pairing in
120 3 Seawater: Composition and Properties

Table 3.7 Ion pairs in seawater—possible pairs and percentage of pairing in standard seawater at
25  C
Cation pairing
Free ions
[%] of
component
Component molality Component C-CO3 C-HCO3 C-SO4 C-Cl
Calcium 85–93 Possible ion CaCO30 Ca Ca(SO4)0 CaCl20,
(Ca) pairs (HCO3)+ CaCl+
0.1–0.3 0.3–1 7–13 0–47
Magnesium 48–90 MgCO30 Mg MgSO40 MgCl+
(Mg) (HCO3)+
0.1–0.3 0.1–1 9–11 0–43
Sodium 83–99 % pairing of NaCO3 NaHCO30 NaSO4 NaCl0
(Na) component 0.002– 0.01–0.1 1–2 0–13
of total 0.01
Potassium 78–99 molality KSO4 KCl0
(K) 1–2 0–18
Anion pairing
Component Free ions Component Ca-A Mg-A Na-A K-A
[%] of
component
molality
Carbonate 9–23 Possible ion CaCO30 MgCO30 NaCO3
CO3 pairs 6–21 44–68 15–21
Bicarbonate 69–81 Ca Mg NaHCO30
HCO3 (HCO3)+ (HCO3)+
2–4 7–19 8–11
Sulphate 39–61 % pairing of Ca MgSO40 NaSO4 KSO4
SO4 component (SO4)0
of total 3–5 12–22 16–37 0.4–1.0
Chloride Cl 83–100 molality CaCl20, MgCl+ NaCl0 KCl0
CaCl+
0–0.9 0–4 0–11 0–0.3

the solution. A distinction is made between ionic strength (Im), at which all the
components of a solution are in the form of free ions, and effective ionic strength (Im,
eff), which is composed of the ionic strength of the fraction of free ions of a solution
and the ion pairs it contains (see Eq. 3.75):
!
1 X
i¼n X
i¼n
I m,eff ¼  mmi,F  Z 2i þ mmi,P  Z 2i ð3:75Þ
2 i¼1 i¼1

Im,eff ¼ effective ionic strength, solvent-based [mol=kgH2 O ]


mmi,F ¼ molality of free component i [mol=kgH2 O ]
mmi,P ¼ molality of paired component i [mol=kgH2 O ]
Zi ¼ charge (valence) of component
3.2 Properties of Seawater 121

Calculation programs used to calculate the speciation of the components of


solutions and their properties, which include ion-pair models, often calculate the
ionic strength as effective ionic strength, without specifically showing this in the
output of the results.
Table 3.9 presents the ion and ion-pair balances for standard seawater at 25  C as
obtained using a calculation program modelling ion association. Table 3.8 gives the
stoichiometric association constants of the various ion pairs used by the software to
calculate ion pairing.
Table 3.9 shows the extent to which the free ions and the components of the
charged ion pairs contribute to calculation of the effective ionic strength. The molal
ionic strength Im of standard seawater, as calculated from all the salt content
components irrespective of ionic pairing, is 0.723 mol/kg H2O. Conversely, the
molal effective ionic strength Im,eff of standard seawater, i.e. with consideration of
ionic pairing, is 0.633 mol/kg H2O with an ionic strength component from the
charged ion pairs of 0.021 mol/kg H2O, i.e. 3.3%.
Of the total molal concentration of the salt content components of standard
seawater, approximately 11% are neutral or charged ion pairs, i.e. around 89% of
the total salt content is free ions. The degree to which, as represented by the ion-pair
model, seawater differs from the behaviour of an ideal solution or from the properties
of freshwater is already noticeable in the case of standard seawater. This tendency is
significantly more pronounced in the case of higher-salinity seawater and
concentrates of the kind produced by seawater desalination.

Table 3.8 Stoichiometric Stoichiometric association coefficient


association constants of ion K CA of standard seawater at 25  C
pairs for standard seawater
Type of ion pair (mol/kg H2O)1
at 25  C
CaCl+ 0.233
CaCl20 0.024
Ca(SO4)0 14.277
Ca(HCO3)+ 5.443
CaCO30 120.882
MgCl+ 1.081
MgSO40 12.610
Mg(HCO3)+ 0.007
MgCO30 73.865
NaCl0 0.294
NaSO4 2.909
NaHCO30 0.496
NaCO3 6.478
KCl0 0.284
KSO4 0.014
122

Table 3.9 Ion pairing and effective ionic strength of standard seawater at 25  C
Molality
Total Free ions Ion pairing Ion pairing rate Zi2*mmi
% of % of
of all of component component % of total free charged
components component molarity uncharged charged total molarity molarity ions ion pairs
Component g-mol=kgH2 O g-mol=kgH2 O % g-mol=kgH2 O % % g-mol=kgH2 O
Calcium Ca 0.0107 0.0090 84.80 0.00104 0.00058 0.00162 15.19 0.14 0.03615 0.00058
Magnesium 0.0528 0.0383 72.47 0.00365 0.01088 0.01453 27.51 1.26 0.15312 0.01088
Mg
Sodium Na 0.4860 0.4424 91.03 0.03390 0.00965 0.04355 8.96 3.76 0.44240 0.00965
Potassium K 0.0106 0.0097 91.54 0.00072 0.00018 0.00089 8.44 0.08 0.00969 0.00018
3

Carbonate 0.0001 0.0000 22.82 0.00006 0.00005 0.00011 77.42 0.01 0.00013 0.00005
CO3
Bicarbonate 0.0017 0.0014 81.25 0.00015 0.00017 0.00032 18.74 0.03 0.00138 0.00017
HCO3
Sulphate SO4 0.0293 0.0149 50.99 0.00456 0.00978 0.01434 49.01 1.24 0.05968 0.00978
Chloride Cl 0.5657 0.5198 91.89 0.03452 0.01130 0.04582 8.10 3.96 0.51980 0.01130
∑ 1.1569 1.0355 0.07860 0.04257 0.12117 10.47 1.22234 0.04257
Ionic strength of free ions Im,F and charged ion pairs Im,P 0.6112 0.0213
Ion strength effective Im,e 0.6325
Data source: PHREEQCi for Windows, Database SIT.dat
Seawater: Composition and Properties
3.2 Properties of Seawater 123

3.2.3.2 Activity Coefficients and Solubility


Reactions of components such as

aX 1 þ bX 2 ⇄cX 3

in electrolyte solutions are determined by the equilibrium relation as follows:

½X 3 c
a b
¼ K eq ð3:76Þ
½X 1   ½X 2 

[Xi] ¼ molar concentration of component Xi [mol/kg, mol=kgH2 O ] ¼ mmi


a, b, c ¼ stoichiometric factor

and its equilibrium constant (Keq). In non-ideal solutions, i.e. in all higher-salinity
solutions such as seawater and the concentrates produced during seawater desalina-
tion, it is not possible to use the analytical molar concentration as presented in
Eq. (3.76). Instead, as already described in Section “Ion Association and Ion
Pairs”, it is necessary to take into consideration the deviation of the solution from
ideal behaviour.

(a) Either the concentration of the respective component Xi is provided with a


correction factor, the so-called activity coefficient (γ X i ), and the thermodynamic
equilibrium constant (K 0eq ) is then used as the equilibrium constant
(b) Or the analytical concentration is left unchanged and K 0eq is replaced by
appropriate, e.g. measured, equilibrium values, which are referred to as apparent
or stoichiometric equilibrium constants (K eq ).

In variant a, activity aX i is then calculated from the molar concentration [Xi] with γ X i
as follows:

aX i ¼ ½X i   γ X i ð3:77Þ

aX i ¼ activity of component Xi [mol/kg, mol=kgH2 O ]


γ X i ¼ activity coefficient of component Xi

and the equilibrium correlation in Eq. (3.76) is then given the form as follows:

γ cX 3  ½X 3 c acX 3
¼ ¼ K 0eq ð3:78Þ
γ aX 1  ½X 1 a  γ bX 2  ½X 2 b aaX 1  abX 2

K 0eq ¼ thermodynamic equilibrium constant


124 3 Seawater: Composition and Properties

With knowledge of the activity coefficients of the individual components and the
stoichiometric equilibrium constant (K eq), it is possible, as in Eq. (3.79), to calculate,
for example, thermodynamic equilibrium constant (K 0eq ) from stoichiometric con-
stant K eq :

γ cX 3
K 0eq ¼ K eq  ð3:79Þ
γ aX 1  γ bX 2

K eq ¼ stoichiometric equilibrium constant

It should therefore be noted that, for thermodynamic modelling of chemical


equilibria or calculation of the solubility of compounds, if activity is used it should
always be used together with K 0eq . If one uses stoichiometric modelling with the
analytical concentrations of the components, then K eq should always be used for this
purpose. The two constants are also differently dependent on the temperature, salt
content, and pressure of the solution. While the thermodynamic equilibrium constant
is dependent on pressure and temperature, the stoichiometric constant is additionally
dependent on the salt content (ionic strength or salinity):

K 0 ¼ K 0 ðT, pÞ ð3:80Þ

K  ¼ K  ðT, p, I Þ ð3:81Þ

The influence of pressure can be left out of consideration when the two constants
are calculated for the pressure range in which seawater desalination treatment
processes are operated.
For the solution reaction of a solid component or precipitation of a precipitate
when the solubility limit of a compound is exceeded according to

X 3S ⇄aX 1 þ bX 2

there applies the equilibrium relation according to

½X 1 a  ½X 2 b
¼ K eq ð3:82Þ
½X 3 cs

The concentration of the solid component or precipitate is not included in the state
of equilibrium of the solution reaction, i.e. the activity of the solid product is set
equal to 1.
Equation (3.82) then changes to Eq. (3.83) and, for formulation as an activity
equation, to Eq. (3.84):
3.2 Properties of Seawater 125

½X 1 a  ½X 2 b ¼ K sp ð3:83Þ

aaX 1  abX 2 ¼ K 0sp ð3:84Þ

K sp ¼ stoichiometric solubility product


K 0sp ¼ thermodynamic solubility product

The dimension of the equilibrium constants and solubility products depends on


the dimension of the component concentrations and the stoichiometric factors of the
corresponding chemical reaction.
The respective value of the stoichiometric or thermodynamic solubility product
(K sp or K 0sp ) characterizes the solubility limit of a low-solubility compound in a
solution. With knowledge of the activity coefficients of the components involved in
the reaction and with knowledge of the value of the stoichiometric solubility product
(K sp ), the thermodynamic solubility product (K 0sp ) can be calculated as

K 0SP ¼ K SP  γaX 1  γbX 2 ð3:85Þ

and vice versa.


If using the stoichiometric calculation method, the product of the dissolved ion
concentrations of a low-solubility compound is known as the ion concentration
product (ICP). If using the thermodynamic calculation method, it is referred to as
the ion activity product (IAP):

½X 1 a  ½X 2 b ¼ ICP ð3:86Þ

aaX 1  abX 2 ¼ IAP ð3:87Þ

ICP ¼ ion concentration product


IAP ¼ ion activity product

In either case, it can be set into relation to the solubility product of the
low-solubility compound. The quotient of this product and the solubility product
of a compound is known as the saturation ratio and denotes the degree of solubility
or saturation of the compound:

IAP ICP
SR ¼ ¼  ð3:88Þ
K 0sp K sp

SR ¼ saturation ratio
126 3 Seawater: Composition and Properties

If the SR value is >1, then the solution is oversaturated with the relevant
compound, which means that the compound can be expected to precipitate. Where
SR is <1, the compound is in solution and, if SR ¼ 1, there is an equilibrium state in
which no precipitation is currently to be expected.
Frequently, the scaling index (SI) is used to characterize the solubility state of a
low-solubility compound Eq. (3.89). This index is the logarithm of the saturation
ratio and is especially suitable for thermodynamic modelling of the solubility state of
a compound:

IAP
SI ¼ log SR ¼ log ¼ log IAP  log K 0sp ¼ log IAP þ pK 0SP ð3:89Þ
K osp

For solution components whose ions form large numbers of ion pairs (particularly
divalent cation and anion components), it may be necessary, especially when
calculating their solubility, in addition to the use of activity coefficients, also to
take into consideration the ion pairing/free ion fraction. The free ion fraction factor
( f F Xi ) of a component Xi is calculated as in Eq. (3.90) or from its stoichiometric
association constant K x using Eqs. (3.71 and 3.72).

½X F i 
f FXi ¼ ð3:90Þ
½X T i 

F X i ¼ free ion fraction factor of Xi

In this case, the activity coefficients of component Xi consist of activity coeffi-


cient (γ FXi ), which is determined from the free ions of the component, and total
activity coefficient (γ T X i ), which is calculated from γ T X i and F X i as in Eq. (3.91).

f FXi  γ FXi ¼ γ T Xi ð3:91Þ

γ FXi ¼ activity coefficient of component Xi based on free ions of Xi


γ T X i ¼ activity coefficient of component Xi based on total ions of Xi

Total activity coefficient (γ T X i ) is then lower by the degree to which component Xi


forms ion pairs. The correspondingly reduced activity of the component is calculated
as follows:

aX i ¼ ½ X F i   γ F X i ¼ ½ X T i   f F X i  γ F X i ¼ ½ X T i   γ T X i ð3:92Þ
3.2 Properties of Seawater 127

[X F i ] ¼ free ion concentration of component Xi


[X T i ] ¼ total ion concentration of component Xi

The extent to which the ion pairing of a given compound needs to be taken into
consideration for an activity coefficient calculated from the analytical concentration
will depend on the extent to which the calculation model used to determine the
activity coefficient already made allowance for the ion pairing and ion speciation of
the solution.
Knowledge of solution behaviour and solubility limits is necessary in seawater
membrane desalination, especially for those low-solubility seawater constituents
that, as they become concentrated during the desalination process, are capable of
forming precipitates (scaling) on the membranes. These are mainly the alkaline earth
sulphates strontium sulphate (SrSO4), barium sulphate (BaSO4), and calcium
sulphate (CaSO4  2H2O) as well as the low-solubility compounds calcium fluoride
(CaF2) and calcium carbonate (CaCO3). The last-named compound is an important
component in the lime/carbonic acid equilibrium of seawater and is discussed in
detail in a later section of this chapter (see Sect. 3.2.4.4).

3.2.3.2.1 Thermodynamic Modelling


Determination of Activity Coefficients To calculate the solubility of these so-called
scale formers or scalants for seawater of a given composition using thermodynamic
modelling, i.e. by means of the activities of the seawater constituents and their
thermodynamic solubility products (K 0CA ), it is necessary to have knowledge of the
activity coefficients of the components of the seawater γCA, their dependence on
temperature and ionic strength or salinity, as well as the respective thermodynamic
solubility product K 0sp and its dependence on temperature.
The activity coefficients can be determined both using the above-described ion
association model [27, 29–31, 33] and the models based on the assumption and
definition of specific interactions between the ions of a solution (specific ion
interaction models). The ion association model can be used to calculate activity
coefficients in solutions with an ionic strength of up to I ¼ approx. 1 m. Beyond that,
specific ion interaction models are better suited.
Specific ion interaction models are based on the Debye-Hückel theory
(DH theory).
For the ions of a concentrated solution, the activity coefficients are normally
smaller than 1. Consequently, the activity of an ion is also lower than its analytical
concentration. This is because the mobility of charged ions in a concentrated solution
decreases on account of the rise in forces and interactions between opposite ionic
charges as the concentration of the solution increases. The Debye-Hückel theory
provides a mathematical definition of this interaction and its dependence on the
concentration or ionic strength of a solution Eq. (3.93–3.93b):
128 3 Seawater: Composition and Properties

pffiffiffiffiffi
log γ Xi ¼ ADH  Z 2Xi  Im ð3:93Þ

1 1:825  106
ADH ¼ ρswðT Þ 2   3 ð3:93aÞ
E ðT Þ  T 2

60, 954
E ðT Þ ≌  68:937 ð3:93bÞ
T þ 116

ADH ¼ 1. Debye-Hückel parameter [kg1/2/mol1/2] ¼ 0.5101 at 25  C


ρSW(T ) ¼ density of seawater at temperature (T ) [kg/dm3]
T ¼ temperature [K] ¼273.15 + t
E(T) ¼ dielectric constant of water (at t ¼ 25  C ¼ 78.38)
Zi ¼ charge (valence) of component

According to the basic version of the Debye-Hückel equation, as derived


directly from the DH theory, the logarithm of the activity coefficient decreases
with the root of the ionic strength and the square of the ionic charge. This equation,
however, is a limiting law, because it is valid only up to an ionic strength of
I < 0.01 ¼ ~ 600 mg/l NaCl.
The range of validity of the Debye-Hückel limiting law equation can be extended
by modifications, with additional ion- and component-specific parameters being
added to the basic equation. Such modifications (see Table 3.10) are the:

• Extended Debye-Hückel equation


• Davies equation
• Truesdell-Jones (WATEQ) equation

Table 3.10 Activity coefficient calculation: Debye-Hückel limiting law, DH equation


modifications, and ion interaction algorithm—range of validity
Range of validity
Ionic strength Salt content as NaCl
Type of algorithm [mol/l, mol/kg] [mg/l NaCl]
Debye-Hückel eq. (DH limiting law) ~ < 0.01 ~ 600
Debye-Hückel limiting law modifications
Extended DH equation ~ < 0.1 ~ 6000
Davies equation ~ 0.1–0.5 ~ 6000–30,000
Truesdell-Jones (WATEQ) equation ~<1 ~ 60,000
Specific ion-interaction algorithms
Brønsted-Guggenheim-Scatchard equation ~ 2–4.5 ~120,000–260,000
(specific interaction theory, SIT)
Bromley equations ~<6 ~ 350,000
Pitzer equations To 6 and above To 350,000 and higher
3.2 Properties of Seawater 129

These modifications allow the range of validity of the Debye-Hückel equation to


be extended to an ionic strength of I ~ 1–58.000 mg/l NaCl, i.e. into the range of
seawater salt content (see Table 3.10).
To make the Debye-Hückel theory applicable to solutions with an even higher
salt content, it is necessary, in addition to ion-specific parameters, also to take into
consideration the interactions in concentrated solutions between the individual types
of charged cations and anions contained in the solution. In the case of highly
concentrated solutions, it is additionally necessary to take account of interactions
between identically charged ions and uncharged ion pairs. To take such interactions
into consideration, the Debye-Hückel equation or its modifications are extended with
so-called secondary and tertiary virial coefficients. This so-called specific interac-
tion theory can then be used to calculate the activity coefficients of more highly
concentrated seawater and concentrates with high salt contents. A large number of
such interaction equations have been developed for general application and also for
specific electrolyte types.
Like the Debye-Hückel equation, the extensions of the interaction equations can
be derived from thermodynamic laws for electrolyte solutions. In some cases,
however, the virial coefficients and interaction parameters must be determined
from or calibrated with empirical measurements. For this reason, interaction
equations are known also as semi-empirical equations.
Table 3.10 gives only those interaction algorithms that have so far been applied in
detail to seawater and seawater concentrates.
These are the:

• Brønsted-Guggenheim-Scatchard equation or SIT


• Bromley equations
• Pitzer equations

However, the range of validity given in Table 3.10 for each of the individual
equations and equation systems should be viewed only as a rough guide. For
so-called 1–1 electrolytes (NaCl, KCl, etc.), it is usually possible for the activity
coefficients to be calculated with sufficient accuracy to a higher ionic strength than is
the case for so-called 2–1 electrolytes (CaCl2, MgCl2), 2–2 electrolytes (BaSO4, Sr
SO4, CaSO4), or electrolytes with an even higher ion configuration. In cases where
the activity coefficient is being calculated mainly for these last-mentioned electrolyte
combinations, it is always recommended to use an equation or equation system with
a higher range of validity. To calculate the scaling in concentrates from desalination
processes, it is advisable to determine the activity coefficients using the Bromley or
Pitzer equations.
In addition to the first Debye-Hückel parameter as in Eq. (3.93), the extended
Debye-Hückel equation Eq. (3.94) contains a second Debye-Hückel parameter and
also an additional coefficient aI:
130 3 Seawater: Composition and Properties

pffiffiffiffiffi
Im
log γ Xi ¼ ADH  Z 2Xi  pffiffiffiffiffi ð3:94Þ
1 þ BDH  ai  I m
1 50:29
BDH ¼ ρswðT Þ 2   12 ð3:94aÞ
E ðT Þ  T

ai ¼ ionic radius [Å] (see Table 3.11)


BDH ¼ second Debye-Hückel parameter [kg1/2/mol1/2] ¼ 0.3284 at 25  C

By extending the basic Debye-Hückel equation with these additions, the algo-
rithm can then be used also for higher concentrations to a maximum ionic strength of
I ¼ 0.1. The second Debye-Hückel parameter is multiplied by coefficient ai. Factor
ai is an empirical parameter with no direct physical equivalent. Its value depends on
the type of ion and can be interpreted as the distance of closest approach of two ions
or also as the apparent mean diameter of the hydrated ion. The extension of the basic
equation as in Eq. (3.94) makes allowance for the fact that ions have a finite radius
and are not exclusively punctiform.
Both Debye-Hückel parameters ADH and BDH are linked to the dielectric constant
of water (ET) and density (ρSW) (see Eqs. 3.93a and 3.94a). According to the
temperature dependence of the dielectric constant Eq. (3.93b) and density
Eqs. (3.39 and 3.40), the two DH parameters then also have a corresponding
temperature dependence (see Fig. 3.26).
Table 3.11 presents values of ion parameter (ai) for different types of cations and
anions.
The Davies equation Eq. (3.95) is the simplest of the extended Debye-Hückel
equations. In addition to the limiting law, it contains an empirical factor b, which is
identical for all types of ions. The activity coefficients of ions calculated using this
equation differ only with regard to their different charge numbers:
 pffiffiffiffiffi 
Im
log γ Xi ¼ Z 2Xi  ADH  pffiffiffiffiffi  b  I m ð3:95Þ
1 þ Im

b ¼ Davies empirical coefficient ¼ 0.2–0.3

Table 3.11 Debye-Hückel Debye-Hückel ion parameter (ai+) [Å]


ion parameters for different
Cation Anion
cations and anions
H3O+ 9 CO32 4.5
Mg 2+
8 HCO3
2+
Ca 6 SO42 4
Ba 2+
5 OH 3.5
Sr 2+
F
Na +
4 NO3 3
K +
3 Cl
NH4 +
Br
3.2 Properties of Seawater 131

Fig. 3.26 Debye-Hückel parameters ADH and BDH—temperature dependence

In addition to the charge number of ions, the Truesdell-Jones equation


Eq. (3.96) additionally takes into consideration ion-specific differences by means
of both the ion parameter ai and an additional ion parameter bi. Table 3.12 gives the
values of these parameters for different types of cations and anions:
pffiffiffiffiffi
Im
log γ Xi ¼ Z 2Xi  ADH  pffiffiffiffiffi þ bi  I m ð3:96Þ
1 þ BDH  ai I m

ai ¼ ion-specific parameter (see Table 3.12)


bi ¼ ion-specific parameter (see Table 3.12)

With this equation, it is sometimes possible to calculate activity coefficients to an


ionic strength of I ~ 1 (i.e. including for a seawater salinity exceeding the standard
composition) with an accuracy similar to that achieved with significantly more
complex interaction equations. This accuracy can be further improved by combining
this equation with the ion association model.
Common to all extended Debye-Hückel equations originating from the specific
interaction theory is that, to calculate the activity coefficient of a specific ion, it is
necessary to take into consideration not only the ion-specific parameters but also the
interactions between all cationic and anionic components of the given solution.
132 3 Seawater: Composition and Properties

Table 3.12 Truesdell- Ion size parameter


Jones equation—ion
Type of ion ai bi
parameters for cations and
anions Cation
H3O+ 4.78 0.24
Mg2+ 5.46 0.22
Ca2+ 4.86 0.15
Ba2+ 4.55 0.09
Sr2+ 5.48 0.11
Na+ 4.32 0.06
K+ 3.71 0.01
Anion
CO32 5.40 0
HCO3
SO42 5.31 0.07
OH 10.65 0.21
F 3.46 0.08
Cl 3.71 0.01

The main difference between the specific ion interaction equations given in
Table 3.10 is the extent to which the algorithms take account not only of the
interactions between the cationic and anionic components of the solution but also
of other interactions between identically charged and uncharged compounds.
The simplest of the ion interaction equations is the Brønsted-Guggenheim-
Scatchard (BGS) equation, known in one of its versions also as the specific
interaction theory. This equation with a second virial coefficient takes into consid-
eration exclusively the interactions between identically charged, strong ions. It
contains an extended Debye-Hückel term by using a constant value (ai ¼ 1 for the
original BGS equation, ai ¼ 1.5 for the SIT equation, depending on the version of the
equation) for ion parameter ai. In its general form, the BGS equation is presented as
follows:
X
log γ Xi ¼ Z 2Xi  D þ ðC $ AÞ ð3:97Þ
pffiffiffiffiffi
Im
D ¼ ADH  pffiffiffiffiffi ð3:98Þ
1 þ BDH  ai  I m

∑(C $ A)¼ cation-anion interactions ¼ (C$Cl) + (C$SO42) + (C$HCO3) + . . .


D ¼ Debye-Hückel term
ADH ¼ first Debye-Hückel parameter
BDHai ¼ 1 in D for BGS equation
¼ 1.5 in D for SIT equation
ZXi ¼ charge (valence) of cationic or anionic component
3.2 Properties of Seawater 133

The activity coefficients for a cation and an anion are calculated as in Eqs. (3.99
and 3.100), respectively. The value of the second virial coefficient is formed by the
sum of all the products of the respective specific interaction coefficient ß(c,a) for each
possible C$A grouping of the specific ion for which the activity coefficient is being
determined and the respective molar concentration of ion grouping C$A:
X
log γ C ¼ Z 2C  D þ βðc,a,I m Þ  ma ð3:99Þ
a
X
log γ A ¼ Z 2A  D þ βðc,a,I m Þ  mc ð3:100Þ
c

ß(c,a) ¼ interaction coefficient for cationic components c with anionic components a


at 25  C [kg/mol]
c ¼ subscript for cations in solution
a ¼ subscript for anions in solution

The range of validity of the BGS/SIT equation can extend to a maximum ionic
strength of I ¼ 4.5. This is especially the case if the background salinity is made up
predominantly of 1–1 electrolytes, such as NaCl. The accuracy of calculation for the
activity coefficients of 2–1 and 2–2 electrolytes can be improved by combining this
equation with the ion association model.
Whitfield used the combination of these two theories to calculate activity
coefficients and osmotic coefficients in seawater [34–37]. He found good agreement
between the results of this equation and those of the Pitzer equations up to an ionic
strength of I ¼ 3 and also with available practical experimentation results. A list of
the interaction coefficients ß(c,a) at 25  C and for different ion configurations similar
to those in seawater can be found in [35]. Additional interaction coefficients, named
εi,k for the SIT version of the equation, are given in [38, 39].
The Bromley equations also contain the Debye-Hückel term with extension by a
virial coefficient. This virial coefficient, however, consists of an interaction coeffi-
cient (Bc,a), which is likewise specific to a certain ion group, and another coefficient
B. Coefficient B is calculated from Bc,a and Im of the solution (see Eq. 3.101b).
Ion-specific activity coefficients for a multi-component electrolyte solution such as
seawater are calculated for a cation and an anion as follows [40, 41]:
X
log γ C ¼ Z 2C  D þ Z c,a 2  ðB þ Bc,a Þ  ma ð3:101Þ
a
X
log γ A ¼ Z 2A  D þ Z c,a 2  ðB þ Ba,c Þ  mc ð3:101aÞ
c

ð0:06 þ 0:6  Bc:a Þ  jZ c  Z a j


B¼  2 ð3:101bÞ
1 þ j1:5I m
Z c Z a j
134 3 Seawater: Composition and Properties

Zc þ Za
Z c,a ¼
2

logγC ¼ activity coefficient of cation


logγA ¼ activity coefficient of anion
BDHai ¼ 1 for Debye-Hückel term D in Bromley equation
Bc,a ¼ interaction coefficient for cationic components c with anionic components at
25  C[kg/mol]

Specific interaction coefficients (Bc,a) as well as cation- and anion-specific


coefficients (Bc and Ba) are given in [39, 40] for a multiplicity of electrolyte
configurations as well as cations and anions at a temperature of 25  C.
Compound-specific coefficients (Bc, a) can be calculated from the cation- and
anion-specific coefficients as in Eq. (3.101c). The required cation and anion
increments are also given in [39, 40].

Bc,a ¼ Bc þ Ba þ δc þ δa ð3:101cÞ

Bc ¼ interaction coefficient for cation [kg/mol]


Ba ¼ interaction coefficient for anion [kg/mol]
δc ¼ interaction increment for cation
δa ¼ interaction increment for anion

Once again, to calculate the activity coefficient of an ion, the value of the second
virial coefficient is formed by the sum of all the products of the respective specific
interaction coefficient (B + Bc,a) and the molar concentration of the respective ion
(ma or mc). The range of validity of the Bromley equations extends to an ionic
strength of I ¼ 6, i.e. far above the salt content of highly concentrated seawater and
the usual concentrates from seawater desalination processes.
In addition to the interactions between strong, differently charged electrolytes, the
Pitzer equations [42–45] take into consideration the interactions between iden-
tically charged ions and with oppositely charged ions as well as between uncharged
components of the kind formed by ion association.
The Pitzer equation for calculating the specific activity coefficient of an ion in a
multi-component solution is presented in its general form
h X X X i
log γ i ¼ Z 2i  Dp þ ðC $ AÞ þ ðC $ CÞ þ ðA $ A Þ
X X 
þ ðC $ C $ A Þ þ CA0 $ CA0 ð3:102Þ

∑(C $ A) ¼ cation-anion interactions ¼ (C$Cl) + (C$SO42) + (C$F) + . . .


∑(C $ C) ¼ cation-cation interactions ¼ (C$Na+) + (C$Mg2+) + (C$Ca2+) + . . .
∑(A $ A) ¼ anion-anion interactions ¼ (A$Cl) + (A$SO42) + (A$F) + . . .
3.2 Properties of Seawater 135

∑(C $ C $ A) ¼ cation-cation-anion and anion-anion-cation interactions ¼


(C$Na+$Cl) + (C$Mg2+$Cl) + (C$Na+ +SO42) + . . .
∑(CA0 $ CA0) ¼ uncharged compound-compound interactions

According to the multiplicity of possible interactions between the solution


components, it is necessary to use a large number of interaction parameters with
which to extend the basic Debye-Hückel equation (see Eqs. 3.103 and 3.103a). Also
the Debye-Hückel-Pitzer function Dp is modified in comparison to the Debye-
Hückel term D in the other ion interaction equations Eqs. (3.103b and 3.103c):

ln γ i ¼ Z 2i  F þ E 1 þ E 2 þ E 3 þ E4 þ E5 ð3:103Þ

F ¼ Dp þ E 6 þ E 7 þ E 8 ð3:103aÞ
 pffiffiffiffiffi 
Im 2  pffiffiffiffiffi
Dp =  A∅  pffiffiffiffiffi þ  ln 1 þ b  I m ð3:103bÞ
1 þ b  Im b
2:303  ADH
A∅ ¼ ð3:103cÞ
3

b ¼ 1.2 at 25  C
Dp ¼ Debye-Hückel-Pitzer function
AØ ¼ Debye-Hückel-Pitzer parameter

To determine the activity coefficient for a specific ion in a multi-component


solution, the interaction parameters of the Pitzer equation in Eq. (3.103) (E1–E5)
and Eq. (3.103a) (E6–E8) are calculated for a cation as in Eq. (3.103d) and for an
anion as in Eq. (3.103e).
E-parameter equations for single cation activity coefficient (γ M):

X
Na
E 1,M = ma 3 ð2 3 BMa þ E  C Ma Þ ð3:103dÞ
a=1
!
X
Nc X
Na
E2,M = mc  2  ΦMc þ ma  ΨMca
c¼1 a¼1

NX
a 21 X
Na
E 3,M = ma  ma0  Ψ a,a0 ,M
a = 1 a0 = aþ1

Nc X
X Na
E 4,M ¼ jZ M j  mc  ma  Cca
c¼1 a¼1
136 3 Seawater: Composition and Properties

X
Nn
E5,M ¼ mn  ð2  λnM Þ
n¼1

Nc X
X Na
E 6,M ¼ mc  ma  B0ca
c¼1 a¼1

X
N c 1 X
Nc
E7,M ¼ mc  mc0  Φ0c,c0
c¼1 c0 ¼cþ1

NX
a 21 X
Nc
E 8,M ¼ ma  ma0  Φ0a,a0
a = 1 a0 = aþ1

E-parameter equations for single anion activity coefficient (γ X):

X
Nc
E 1,X = mc 3 ð2 3 BcX þ E  C cX Þ ð3:103eÞ
c=1
!
X
Na X
Nc
E 2,X = ma  2  ΦXa þ mc  ΨXac
a¼1 c¼1

NX
c 21 X
Nc
E 3,X = mc  mc0  Ψ c,c0 ,X
c = 1 c0 = cþ1

Nc X
X Na
E 4,X ¼ jZ X j  mc  ma  Cca
c¼1 a¼1

X
Nn
E 5,X ¼ mn  ð2  λnX Þ
n¼1

Nc X
X Na
E6,X ¼ mc  ma  B0ca
c¼1 a¼1

X
N c 1 X
Nc
E 7,X ¼ mc  mc0  Φ0c,c0
c¼1 c0 ¼cþ1

NX
a 21 X
Nc
E 8,X ¼ ma  ma0  Φ0a,a0
a=1 a0 = aþ1
X
E¼ mi  Z i
i
3.2 Properties of Seawater 137

mc,a,n,i ¼ molality of cationic or anionic or neutral or individual compound [g-mol/


kg]
c ¼ subscript for cation in solution
a ¼ subscript for anion in solution
n ¼ subscript for neutrals in solution
i ¼ subscript for compounds in solution ¼ c or a or n
M ¼ subscript for cation of interest
X ¼ subscript for anion of interest
Nc ¼ number of cation
Na ¼ number of anion
Nn ¼ number of neutral
Zi ¼ charge (valence) of cationic or anionic component i
mi ¼ molality of cationic or anionic component I [mol/kgsolvent]
Bca, B0ca ¼ parameters for binary ion interaction of mono- and divalent cation and
anion
Cca ¼ parameter for triple ion interaction of mono- and divalent cation and anion
Φca, Ψca ¼ parameters for binary like-sign cation-cation and anion-anion
interactions
λca ¼ parameter for neutral solute interactions

The parameters for the binary interaction between differently charged cations and
anions (Bca and B0ca) are, in turn, dependent on the ionic strength (Im) of the solution
and are calculated from the basic parameters βðca0Þ , βðca1Þ , and βðca2Þ as in Eqs. (3.103f
or 3.103g).
Equation 3.103f applies to the calculation of electrolytes of configuration 1–1
(e.g. NaCl), 1–2 (e.g. Na2SO4), and 2–1 (e.g. CaCl2). Equation 3.103g is used for
higher-valence electrolytes (e.g. 2–2 electrolytes such as CaSO4, BaSO4, SrSO4). In
this way, Pitzer takes into consideration the greater tendency to ion pairing of 2–2
electrolytes and higher-valence ions [44].
 pffiffiffiffiffi
Bca ¼ βðca0Þ þ βðca1Þ  g  α  I m ð3:103fÞ
pffiffiffiffiffi
0 ð1Þ 0 α  Im
Bca ¼ βca  g 
Im
2  ½1  ð1 þ xÞ  ex

x2
 
0 2  1  1 þ 12 x2  ex
g ¼
x2
pffiffiffiffiffi
x ¼ α  Im

α¼2
138 3 Seawater: Composition and Properties

 pffiffiffiffiffi  pffiffiffiffiffi
Bca ¼ βðca0Þ þ βðca1Þ  g  α1  I m þ βðca2Þ  g  α2  I m ð3:103gÞ
pffiffiffiffiffi pffiffiffiffiffi
0 ð1Þ 0 α1  I m ð2Þ 0 α2  I m
Bca ¼ βca  g  þ βca  g 
Im Im
α1 ¼ 1:4

α2 ¼ 12:0

Parameter Cca for the triple reaction between mono- and divalent cations and
anions is dependent
  on the charge number of the ions and is calculated from basic
parameter C ∅ ca as in Eq. (3.103h). Conversely, parameter (Φca) for the reaction
between identically charged ions is dependent on the ionic strength of the relevant
basic parameter E θca [46] Eq. (3.103i).

C∅
ca
C ca ¼ 1 ð3:103hÞ
2  jZ c  Z a j2

Φca ¼ θca þ E θcaðI m Þ ð3:103iÞ

Φ0ca ¼ E θcaðI m Þ

Parameter (λca) for the interaction of neutral compounds needs to be included in


the calculations only if, for example, the gas solubility (e.g. of carbon dioxide) or
neutral ion pairs, such as CaSO40 and MgSO40, are of importance for calculating the
activity coefficient. Otherwise this interaction parameter can be left out of
consideration.
To calculate the activity coefficient for an ion in a multi-electrolyte solution such
as seawater, it is necessary to have knowledge of all the basic parameters βðca0Þ , βðca1Þ

, βðca2Þ , C ∅ E
ca , θ ca , θca , and Ψca. Additionally required is a full analysis of the solution
from which it is possible to calculate the molality of the individual ions and,
therefrom, the ionic strength of the solution.
The basic Pitzer parameters for ions and compounds for a temperature of 25  C
are documented in a large number of publications [39, 43–45]. For seawater-like
electrolyte solutions as well as for seawater itself, Harvie and colleagues have
published a number of works [46, 47] with data on Pitzer parameters that form the
basis for the calculation of seawater properties using the Pitzer equations. However,
the parameters for BaSO4 and SrSO4 are not included in these data. For these Pitzer
parameters, reference is made to [48–50].
The primary interaction parameters of the Pitzer equations are temperature-
dependent as in Eq. (3.103j) for temperatures above 25  C. However, the coefficients
of the polynomial are different for every compound and its interaction parameters.
They are given in [51, 52] for various compounds and relevant Pitzer parameters.
3.2 Properties of Seawater 139

Fig. 3.27 Single-ion activity coefficients of seawater of constant proportions—dependence on


ionic strength (Data source: PHREEQCi, Database Pitzer.dat)

Also for the other above-given semi-empirical interaction equations (BGS/SIT and
Bromley equations), the interaction parameters are temperature-dependent.

a3 a5 a7
PðT Þ ¼ a1  a2  T þ þ a4  ln T þ þ a6  T 2 þ
T T  263 680  T
a8
þ ð3:103jÞ
T  227

ð0Þ ð1Þ
P(T ) ¼ parameter at temperature (T ), where P(T ) ¼ βcaðT Þ , βcaðT Þ , C ∅
caðT Þ , θ caðT Þ , ΨcaðT Þ:

Figure 3.27 presents values of activity coefficients and their dependence on ionic
strength for the main ions contained in seawater. The calculated data apply to
seawater of constant proportions.
The graph in Fig. 3.28 shows the influence of temperature in the range of 0–40  C
on the values of the activity coefficients for various seawater ions. The graph clearly
shows that, at temperatures of 10–40  C, at which most membrane seawater desali-
nation plants are operated, the influence of temperature on the activity of the
seawater constituents is negligibly small.
140 3 Seawater: Composition and Properties

Fig. 3.28 Single-ion activity coefficients of standard seawater—dependence on temperature (data


source: PHREEQCi, Database Pitzer.dat)

If interaction equations are used to calculate the properties of multi-electrolyte


solutions such as seawater and the concentrates from desalination processes, this
calls for highly complex computational processes, it being necessary to take into
consideration a multiplicity of input and calculation parameters. This makes manual
calculation virtually impossible, for which reason suitable calculation programs or
software modules are required. More details can be found in the section on “Soft-
ware for Thermodynamic Solution Modelling”.

Determination of the Thermodynamic Solubility Product The thermodynamic


solubility product (K 0sp ) of a low-solubility substance can be calculated from the
standard Gibbs free energy of reaction (ΔG0r Þ of its formation from the reactants as
follows:

ΔG0r
ln K 0sp ¼  ð3:104Þ
RT
ΔG0r
log K 0sp ¼  ð3:104aÞ
5:70801
3.2 Properties of Seawater 141

K 0sp ¼ thermodynamic solubility product


ΔG0r ¼ standard Gibbs free energy of reaction [kJ/mol]
R ¼ universal gas constant ¼ 8.314462 103[kJ/mol, K]

This, in turn, is calculatedfrom thestandard Gibbs free energy of formation of the


individual reacting products ΔG0f ,P and reactants (ΔG0f ,R Þ in consideration of the
respective stoichiometric factors of the participating components as follows:
X X
ΔG0r ¼ νp  ΔG0f ,P  νR  ΔG0f ,R ð3:105Þ

ΔG0f ,P ¼ standard Gibbs free energy of formation of products [kJ/mol]


ΔG0f ,R ¼ standard Gibbs free energy of formation of reactants [kJ/mol]
νp ¼ stoichiometric factor of product
νR ¼ stoichiometric factor of reactant

The standard Gibbs free energy of formation of the individual ions and
compounds can normally be found in tables and databases. Where it is unavailable
for a given compound, it can be calculated as in Eq. (3.106) from the standard
enthalpy of formation (ΔH 0f Þ of the component and its standard entropy of formation
 
ΔS0f .

ΔG0f ¼ ΔH 0f  T  ΔS0f ð3:106Þ

ΔH 0f ¼ standard enthalpy of formation at 298.15  C [kJ/mol]


ΔS0f ¼ standard entropy of formation at 298.15  C [kJ/mol, K]

Table 3.13 summarizes the thermodynamic data for calculating the standard
Gibbs free energy of reaction and, therefrom, the solubility products of the most
important low-solubility compounds in seawater responsible for scaling in seawater
desalination plants.
Using Eqs. (3.104a and 3.105) and the thermodynamic data from Table 3.13, the
solubility product at 25  C and its pK 025 C value are calculated for strontium sulphate
SrSO4 according to the following example:

SrSO4 ⇄Sr2þ þ SO2


4

ΔG0r ¼ ΔG0Sr2þ þ ΔG0SO2  ΔG0SrSO4


4

ΔG0r ¼ ð558:66Þ þ ð744:54Þ  ð1340:97Þ

ΔG0r ¼ 37:77 ½kJ=mol


142 3 Seawater: Composition and Properties

Table 3.13 Thermodynamic data of selected seawater compoundsa


Standard enthalpy of Standard Gibbs energy Standard entropy of
formation of formation formation
ΔH 0f ΔG0f ΔS0f
Compound/ion kJ/mol kJ/mol J/mol, K
Ba2+ 537.64 560.74 +9.6
Ca2+ 543.0 553.54 56.2
Mg2+ 467.0 454.8 137.0
Sr2+ 545.8 559.44 32.6
HCO3 689.93 586.90 +98.4
CO32 675.23 527.90 50.0
F 335.35 278.8 13.8
SO42+ 909.34 744.54 +18.5
BaSO4 1473.19 1362.2 +132.2
SrSO4 1453.10 1340.97 +117.0
CaSO4  2H2O 2022.6 1797.5 +194.1
CaF2 1228.0 1175.6 +68.6
CaCO3 (calcite) 1207.6 1129.1 +91.7
CaCO3 1207.8 1128.2 +88.0
(aragonite)
Ca[Mg(CO3)2 2326.3 2163.6 +155.8
(dolomite)
Mg(OH)2 924.7 833.7 +63.24
H2O 285.83 273.12 +69.90
H+ 0 0 0
OH 230.02 157.24 10.90
a
CODATA Key Values for Thermodynamics
CRC Handbook of Chemistry and Physics, CRC Press; [48–50]

37:77
log K 0SrSO4,25 C ¼  ¼ 6:617
5:70801
pK 0SrSO4,25 C ¼ 6:617

To determine the solubility product for other temperatures, it is possible to make a


conversion using Eq. (3.107) and the standard enthalpy of formation ΔH 0f of the
products and reactants from Table 3.13. The enthalpy of reaction ΔH 0r is then
calculated in the same way as shown in the above example for determining the
standard Gibbs free energy of reaction for strontium sulphate using ΔH 0f instead of
ΔG0f .
 
ΔH 0r 1 1
ln K 0sp,T 2 ¼ ln K 0sp þ   ð3:107Þ
R 298:15 T 2
3.2 Properties of Seawater 143

R ¼ 8.314462 103 [kJ/mol, K]


ΔH 0r ¼ standard enthalpy of reaction [kJ/mol]

The thermodynamic data given in Table 3.13 are temperature-dependent. In the


temperature range of 10–50  C, in which membrane desalination plants normally
operate, the influence of temperature is only small and can be ignored for manual
calculation of the thermodynamic solubility product. For thermal desalination pro-
cesses with operating temperatures above 100  C, however, it is necessary to take
into consideration the temperature dependence of the thermodynamic parameters.
This influence of temperature, however, is already taken into account in the follow-
ing polynomial equations Eqs. (3.108 and 3.109).
For barium sulphate (BaSO4), calcium sulphate (CaSO4  H2O), and strontium
sulphate (SrSO4), the thermodynamic solubility product can be calculated as a
function of temperature using a polynomial Eq. (3.108) and the corresponding
coefficients [48–50]:

A  ln T B  T C I Ig
ln K 0sp ¼ þ þ  h  ð3:108Þ
R 2  R 2  R  T2 R  T R

R ¼ 8.314462 [J/mol, K]

Compound Formula A B C Ih Ig
Barite BaSO4 +594.534 1.91171 40.0731106 200,488 +3740.12
Celestite SrSO4 +641.541 1.90146 42.7605106 251,748 +4102.24
Gypsum CaSO4  +763.714 2.04731 43.2002106 282,176 +4837.58
2H2O

The solubility product of calcium fluoride (CaF2) can be determined as follows [53]:

3120:98 2088:47
log K 0sp ¼ 109:25 þ 0:0024  T   37:63  log T 
T T2
298:4
 4:9  107  T 2  pffiffiffiffi ð3:109Þ
T
The pK 0sp values for the above-mentioned low-solubility compounds in seawater
and their dependence on temperature are presented in a graph in Fig. 3.29.
With the activity coefficients presented in the graph in Fig. 3.27, it is possible, as
in Eq. (3.77), to calculate the activity of seawater components as molality
[mol=kgH2 O ] on the basis of their analytical concentrations. As in Eq. (3.87), one
then obtains from these values the ion activity product (IAP). As in Eq. (3.89) and
the pK 0sp values shown in the graph in Fig. 3.29, this then yields the scaling potential
index (SPI), which characterizes the degree of saturation of the respective compound
in the seawater or concentrate.
144 3 Seawater: Composition and Properties

Fig. 3.29 Thermodynamic solubility products of alkaline earth sulphate compounds—dependence


on temperature

When solubility products are calculated on a thermodynamic basis, however, it is


necessary to bear in mind that the obtained results are fully valid only for pure
substances. When low-solubility mixed products are formed, their solubility limits,
with reference to the individual substances, are subject to significant change,
depending on the solubility of the included foreign compound, with an ensuing
reduction or increase of the solubility of the mixed product.
Magnesium has a strong tendency to combine with sulphate to form the neutral
ion pair MgSO40 (see Tables 3.7 and 3.8), which reduces the sulphate concentration
available for precipitation of the other low-solubility alkaline earth sulphates BaSO4,
SrSO4, and CaSO4  2H2O. In the case of solutions with a higher magnesium
content, such as the concentrates from seawater desalination processes, the scaling
potential index (SPI) should be calculated on the basis of the concentration of
available sulphate as in Eq. (3.73) and not on the basis of the analytical sulphate
concentration.
3.2 Properties of Seawater 145

Software for Thermodynamic Modelling Calculation programs for modelling


characteristic properties of solutions, such as the degree of ion association of their
components, their activity coefficients and solubility products, and the osmotic
coefficient of the solution, contain data sets with the necessary calculation
parameters, such as:

• Ion association constants


• Ion parameters for the Debye-Hückel equations
• Interaction coefficients for interaction equations and polynomial equations with
the relevant coefficients for their temperature dependence
• Thermodynamic parameters for calculating the thermodynamic equilibrium
coefficients and solubility products, including polynomials for calculating their
temperature dependence

The required input data are:

• As complete and consistent an analysis as possible of all the solution components


• Temperature and possibly also pressure of the solution under process conditions

According to the given solution composition, the calculation routines of the


program then assign the corresponding calculation parameters from the data sets
and calculate and output the solution parameters. The quality of the output data
depends on the scope and accuracy of the data available in the data sets for the
individual solution parameters.
Especially in the field of geochemistry, calculation programs have been devel-
oped for modelling the properties of solutions and how they change in the case of
precipitation of low-solubility components, adsorption, and ion exchange and also
for the simulation of transport phenomena. The same calculation programs can be
used to calculate the above-described parameters in seawater. An example of such
software is PHREEQC, a calculation program that was developed for the US
Geological Survey (USGS) of the US Department of the Interior and which is also
maintained and updated by that institution. The program is available free of charge
and various versions of the software can be downloaded from the homepage (http://
wwwbrr.cr.usgs.gov/projects/GWC_coupled/phreeqci/) or from the links given
there. Libraries are available for integrating the PHREEQC algorithms into calcula-
tion programs programmed with C++, C, or FORTRAN. There is also a Microsoft
COM module with which PHREEQC can be integrated into applications with Visual
Basic and Excel.
The same homepage also provides access to extensive documentation on the
individual program versions [54–56]. A detailed introduction to using PHREEQC,
particularly with reference to the modelling of seawater, is given in [57].
The program contains a total of nine databases, each of which includes various
data sets with the necessary parameters for speciation modelling. Each database can
be preselected for entering data in the model. Depending on which database has been
selected, the activity coefficients are calculated either with variants of the Debye-
146 3 Seawater: Composition and Properties

Hückel equation or with routines that use interaction equations (SIT or Pitzer). The
output data from the program differ depending on which database was selected.
The databases phreeqc.dat, llnl.dat, minteq.dat, minteq.v4.dat, and wateq4f.dat
calculate the activity coefficients using the extended Debye-Hückel equation, the
Truesdell-Jones equation, or—if the corresponding ion parameters for the entered
components are unavailable—the Davies equation. These databases additionally
include a speciation calculation using the ion association theory.
If the sit.dat database is selected, the calculation is performed using the SIT
interaction equation, likewise coupled with ion association algorithms. The pitzer.
dat database uses exclusively the Pitzer equations for simulation.
For speciation calculation for a given water composition, the data output from
PHREEQC is made up of:

• Specific data of the solution, such as its ionic strength and, in the case of databases
with interaction equations, the osmotic coefficient of the solution
• List of the component types present in the solution with their respective molality,
activity coefficients, and activity
• List of the solubilities of compounds, each with the logarithm of the solubility
product and ion activity product as well as the saturation index/saturation poten-
tial index

The output data of the various databases can be differentiated according to the
algorithm assigned to the database and also according to the contents of the respec-
tive thermodynamic data sets. For example, with those databases that use ion
association algorithms (all databases with the exception of pitzer.dat), the ionic
strength of the solution is output as the effective ionic strength (Im,eff) (see
Eq. 3.75 and Table 3.9). Moreover, the effective ionic strengths output by the
individual databases may differ depending on which ion association constants are
assigned to the compounds in the data sets. For calculations with pitzer.dat, the ionic
strength is determined from the analytical concentrations of the components without
consideration of ion pairing (output of Im as in Eq. 3.63).
For speciation simulation as well as for determination of the saturation concen-
tration and scaling tendency of seawater and concentrates from seawater desalination
processes, the Pitzer database pitzer.dat should be used for calculation with
PHREEQC.

3.2.3.2.2 Stoichiometric Modelling


To determine the saturation potential of low-solubility compounds with the stoichio-
metric solubility product (K sp ), its value must be known for the relevant calculation
conditions as a function of salt content or ionic strength as well as temperature and
pressure. Also required is the analytical molal concentration of the compound under
investigation (see Eqs. 3.81 and 3.88).
The solubility product and its dependence on the salt content of the solution and
the ambient conditions can be:
3.2 Properties of Seawater 147

(a) Determined by solution tests with corresponding natural or synthetic solutions


and additions of the corresponding compounds in solid form. Strictly speaking,
however, the thus obtained results and their mathematical evaluation are rele-
vant only for the test conditions (composition and salt content of the solution,
test temperature, and test pressure) at which they were determined. If the
obtained algorithms are applied to other solution conditions, then one must
expect inaccuracies, which will be all the greater, the more the effective
conditions differ from the test conditions. In oceanography, as such tests are
normally carried out with synthetic seawater or even solutions with simple dual
or tertiary electrolyte combinations and usually only to a salinity equivalent to
40 g/kg, the application of such results to concentrates from seawater desalina-
tion processes must be viewed with caution.
(b) Derived from the algorithm for the relevant thermodynamic solubility product
and the associated activity coefficients as in Eq. (3.85). The results of a thus
obtained mathematical formula can then be checked for consistency against
existing test data and within the range of validity of those data, and the
algorithms can be suitably adapted. The thus obtained polynomials can then
be better extrapolated and applied to solutions and pressure/temperature
conditions outside of the range of available test data.

If usable algorithms are available for the stoichiometric solubility product, this
method of determining the solubility limits of low-solubility compounds is consid-
erably simpler than thermodynamic modelling. This is also why this method of
calculation is preferred for determining the scaling potential in membrane seawater
desalination plants (see Sects. 5.3.4 and 5.3.4.1 for details).
Equation 3.110 and the relevant coefficients for barite (BaSO4) and celestite
(SrSO4) were derived using the above method b [58].

C
ln K sp,T,S ¼ A þ B  ln T þ þ D  Sn ð3:110Þ
T

Compound Formula A B C D n
Barite BaSO4 +247.616 38.3326 15,421.2 +1.2645 0.3
Celestite SrSO4 +259.542 40.6673 12,691.4 +1.1832
Strontium barite +247.88 38.333 15,421.20 +1.265

Figures 3.30 and 3.31 show the dependence, calculated according to the above
equation, of the pK sp value on salinity and temperature for barium sulphate and
strontium sulphate, respectively.
148 3 Seawater: Composition and Properties

Fig. 3.30 Stoichiometric solubility product of barium sulphate—dependence on temperature and


salinity

As already mentioned, the solubility behaviour of a pure compound changes if


foreign ions are included in the solid product when it is precipitated. Figure 3.32
presents the change in the stoichiometric solubility product of barium sulphate if,
during precipitation, it is present in the form of the mixed substance strontium barite,
i.e. with a concentration of approx. 13 mol% strontium in it. Through inclusion of
the more highly soluble strontium, the solubility of the strontium barite increases as
compared with pure barium sulphate, i.e. the pK sp value of the mixed compound
exhibits a corresponding decrease.

3.2.3.3 Osmotic Pressure and Osmotic Coefficient


From a thermodynamic standpoint, the solution of a substance in a solvent has a
lower chemical potential or less free energy than the solvent itself. If the solution is
separated from the solvent by a semipermeable membrane that allows only
molecules of the solvent to pass, then solvent will flow from the solvent zone to
the solution owing to the difference in chemical potential between the two liquids.
The osmotic pressure is the pressure that must be exerted on the solution side in
order to prevent the transport of solvent through the membrane. From the equilib-
rium relation with the difference in chemical potential between the two liquids and
the osmotic pressure on the solution side, it is possible, at a thermodynamic level, to
derive a relation for calculating the osmotic pressure:
3.2 Properties of Seawater 149

Fig. 3.31 Stoichiometric solubility product of strontium sulphate—dependence on temperature


and salinity

RT
π¼  ln ðaW Þ ð3:111Þ
V W,swðT,I m Þ

π ¼ osmotic pressure [bar; atm]


Vw,sw(T,Im) ¼ molar volume of water in seawater [l/mol]
R ¼ universal gas constant ¼ 8.314462 102 [lbar/mol, K]
¼ 8.205746 102 [latm/mol, K]
aw ¼ water activity

For seawater as a concentrated solution, the osmotic coefficient ϕ, similar to the


activity coefficient described in Sect. 3.2.3.2, characterizes the deviation of the
seawater from the behaviour of an ideal solution:

ln ða Þ  1000
∅ ¼ P w ð3:112Þ
mm,i  M H2 O
i
150 3 Seawater: Composition and Properties

Fig. 3.32 Stoichiometric solubility product comparison—pure BaSO4 and barium-strontium


mixed compound


P ¼ osmotic coefficient of solution P P P
mmi ¼ molality of all compounds in solution ¼ mm,c þ mm,a þ mm,n
i c a n

[mol=kgH2 O ]
MH2 O ¼ molecular weight of water [g/mol]

The parameters osmotic pressure π and osmotic coefficient ∅ can be combined by


water activity aw.
From the above-presented definition of the equilibrium state of a saline solution
and a solvent (water) across a semipermeable membrane as well as from the relation
for the osmotic coefficient, it is possible to derive the following Eq. (3.113) for the
osmotic pressure of a non-ideal solution:

R  T  M H2 O X
π¼ ∅ mm,i ð3:113Þ
V W,swðT,I m Þ  1000 i

The molar volume of water in seawater VW,sw is dependent on the temperature (t)
and ionic strength (Im) of the seawater and can be calculated from the molar volume
of water (Vw,0) as follows:
3.2 Properties of Seawater 151

2

V W,swðtÞ ¼ 4V W,0,25 C  103  2:654  103  7:360  105  t þ 9:171

3
 I 3m=2 5
107  t 2  2:718  109  t 3   103
1:2457

ð3:113aÞ
M H2 O
V W,0 ¼ ð3:113bÞ
ρW ð T Þ

M H2 O ¼ molecular weight of water [g/mol] ¼ 18.01528


Vw,0 ¼ molar volume of water ¼ 18.0686 103 [l/mol] at 25  C
ρW ¼ density of pure water ¼ 997.048 [g/l] at 25  C

From the molecular weight of water (M H2 O ) and the molar volume of seawater
(VW,sw), it is possible to derive a correction factor ( fvw) for how the osmotic pressure
is influenced by the molar volume of water in seawater and its dependence on
temperature and ionic strength:

M H2 O
f vw ¼ ð3:113cÞ
V W,swðT,I m Þ  1000

fvw ¼ factor for molar volume of water in seawater

The equation for osmotic pressure π then has the form


X
π ¼ R  T  f vw  ∅  mmi ð3:114Þ
i

m ¼ molality ¼ mol per kg of solvent (water, H2O) [mol=kgH2 O ]


Pm P P P
mmi ¼ molality of all compounds in solution ¼ mm,c þ mm,a þ mm,n
i c a n

[mol=kgH2 O ]

However, factor ( fvw) is of an order of magnitude of only 0.997 and changes with
the temperature and ionic strength of the seawater by only around three units in the
fourth decimal place. Consequently, the influence of the molar volume of water in
seawater and its dependence on temperature and ionic strength can be neglected for
practical calculations of osmotic pressure. Equations (3.113 and 3.114) can then be
152 3 Seawater: Composition and Properties

simplified to Eq. (3.115), which without osmotic coefficient ϕ is known also as the
van ‘t Hoff equation of osmotic pressure for diluted solutions:
X X
π ¼RT ∅ mmi ¼ R  ð273:15 þ t Þ  ∅  mm,i ð3:115Þ
i i

To calculate the osmotic pressure using Eq. (3.115), it is necessary to enter the
sum of the concentrations of the solution components as molality. Modified calcula-
tion equations for the use of other units of concentration are as follows;

1 1000 X
π ¼ R  ð273:15 þ t Þ   ∅ mMv,i ð3:115aÞ
ρsw 1000  S i

mMv,i ¼ molarity of component i per volume ¼ mol per liter of solution [mol/l]

1000 X
π ¼ R  ð273:15 þ t Þ  ∅ mMm,i ð3:115bÞ
1000  S i

mMm,i ¼ molarity of component i per mass ¼ mol per kg of solution [mol/kg]


X cm H
2 O,i
π ¼ R  ð273:15 þ t Þ  ∅  ð3:115cÞ
i
Mi

M ¼ molecular weight of component i [g/mol]


cmH2 O,i ¼ mass of component i per solvent (H2O) [g/kgH2O]
1000 X cm,i
π ¼ R  ð273:15 þ t Þ  ∅ ð3:115dÞ
1000  S i
Mi

cm,i ¼ Concentration of component i per mass of solution [g/kg]

1 1000 X cv,i
π ¼ R  ð273:15 þ t Þ   ∅ ð3:115eÞ
ρsw 1000  S i
Mi

cv,i ¼ concentration of component per volume of solution [g/l]

1 1000 X DSv,i
π ¼ R  ð273:15 þ t Þ   ∅ ð3:115fÞ
ρsw 1000  S i
M i  1000

DSv,i ¼ concentration of component i of dissolved solids per volume of solution


[mg/l]
3.2 Properties of Seawater 153

1000 X DSm,i
π ¼ R  ð273:15 þ t Þ  ∅ ð3:115gÞ
1000  S i
M i  1000

DSm,i ¼ concentration of component i of dissolved solids per mass of solution


[mg/kg] ¼ [ppm]

For seawater desalination by membrane processes, the osmotic pressure is an


important parameter for process calculation. In order to achieve water transport
through a reverse osmosis membrane, it is necessary to apply a pressure to the
membrane that is higher than the osmotic pressure of the seawater. Therefore, the
osmotic pressure is determining not only for the transport performance of a reverse
osmosis membrane but also for the energy consumption of the desalination process.
Osmotic coefficient Ø is thermodynamically linked with activity coefficient γ by
the Gibbs-Duhem equation, and each of the two parameters can be calculated from
the other by means of integrals derived from the equation.
Therefore, in addition to the algorithms for the calculation of activity coefficients,
the equations of the specific interaction theory also contain algorithms for determin-
ing the osmotic coefficient. The equations of Bromley and Pitzer, and the therefrom
derived algorithms, are most frequently used to calculate osmotic coefficients.
To calculate the osmotic coefficient, the Bromley equation uses an interaction
coefficient B, which is calculated for seawater as described in [41] and as follows:
P
mm,i  Z 2i pffiffiffiffiffi
i I
∅ ¼ 1  ln ð10Þ  ADH  P  m  σ þ ln ð10Þ
mm,i 3
i
2 P 3
mm,i  Z 2i
6 i I I n 7
 4ð0:06 þ 0:6  BÞ  P  mψ þB mþ  C  Imn5
mm,i 2 2 nþ1
i

ð3:116Þ

3 pffiffiffiffiffi 1  pffiffiffiffiffi
σ¼  1 þ Im  pffiffiffiffiffi  2  ln 1 þ I m
I m 3=2 1 þ Im
" #
2 1 þ 2  a  I m ln ð1 þ a  I m Þ
ψ¼  
a  Im ð1 þ a  I m Þ2 a  Im
P
mm,i Z i
2

jZ þ  Z  j ¼ i Pm ffi 1:2457 for seawater


m,i
i

P
1:5  mm,i
1:5 i 1:5
a¼ ¼ P ffi ffi 1:20414
jZ þ  Z  j mm,i  Z 2i 1:2457
i
154 3 Seawater: Composition and Properties

T  243 B1
B ¼ B0  ln þ þ B2 þ B3  ln ðT Þ
T T
C1
C¼ þ C2
T
n ¼ 1:5

B0 ¼ +0.031102 B2 ¼ +1.98989
B1 ¼ 92.220 B3 ¼ 0.27491
C1 ¼ +11.05 C2 ¼ 0.03782

Sharqawy M.H et al. have published in [17] a polynomial for calculating the
osmotic coefficient correlated from calculation data of the Bromley equation. These
data are compiled in [41]:

∅ ¼ a1 þ a2  t þ a3  t 2 þ a4  t 4 þ a 5  S þ a6  S  t þ a 7  S
 t 3 þ a8  S2 þ a9  S2  t þ a10  S2  t 2 ð3:116aÞ

a1 ¼ +8.9453232101 a4 ¼ +2.221121011 a7 ¼ 1.3526261011


a2 ¼ +4.1561074104 a5 ¼ 1.144546104 a8 ¼ +7.0132356106
a3 ¼ 4.6262121106 a6 ¼ 1.478346106 a9 ¼ +5.69605108
a10 ¼ 2.8624031010

S ¼ salinity [g/kg]
Range of validity of Eq. (3.116a): S ¼ 10–120 g/kg; t ¼ 0–200  C

Similar to calculation of the activity coefficient, the Pitzer equation for the
osmotic coefficient [45, 46] takes into consideration the interactions of the seawater
components with regard to both differently and identically charged ions. To calcu-
late ∅, therefore, the same basic parameters βðca0Þ , βðca1Þ , βðca2Þ , C ∅ E
ca , θ ca , θ ca , and Ψca
are used for determining the activity coefficient. However, the coefficients in the
individual equation groups are in some cases calculated differently than in the
activity coefficient equation:
3=2
!

2 A  I m
∅¼1þP   pffiffiffiffiffi þ O1 þ O2 þ O3 ð3:117Þ
mm,i 1 þ b  Im
i

Nc X
X Na
 
O1 ¼ mc  ma  BΦ
ca þ E  C ca
c¼1 a¼1
!
NX
c 1 X
Nc X
Na
O2 ¼ m c  m c0  Φ∅
cc0 þ ma  Ψ cc0 a
c¼1 c0 ¼cþ1 a¼1
3.2 Properties of Seawater 155

!
NX
a 1 X
Nc X
Nc
O3 ¼ ma  ma0  Φ∅
aa0 þ mc  Ψ aa0 c
a¼1 a0 ¼aþ1 c¼1

A∅ = 2:303ADH
3 ; b ¼ 1.2 at 25 C;

1=2 1=2
BΦ ð0Þ ð1Þ
ca ¼ βca þ β ca  e
α1 I m
þ βðca2Þ  eα2 I m

C∅
ca
C ca ¼ 1
2  jZ c  Z a j2
E
Φ∅ E 0
ca ¼ θ ca þ θ caðI m Þ þ I m  θ caðI m Þ
X
E¼ mi  Z i
i


ca ¼ parameter for binary ion interaction of mono- and divalent cation and anion
Cca ¼ parameter for triple ion interaction of mono- and divalent cation and anion
Φ∅
ca , Ψca ¼ parameters for binary like-sign cation-cation and anion-anion
interactions

With the PHREEQC calculation program, after selection of the data set pitzer.dat
and input of the temperature and a seawater analysis, the osmotic coefficient of the
seawater can be calculated under these conditions.
Another calculation equation for the osmotic coefficient was published in [59] by
the former membrane manufacturer DuPont for calculation of its PERMASEP
hollow-fibre membranes:

S 1
∅¼1  T DH  2  ln ðT DH Þ  þ B  I0 þ C
3:375  I m T DH
 ðI 0 Þ
2
ð3:118Þ
0P 1ρ
mm,i  Z 2i W ðT Þ  3=
B i C 2:3375556  104 2
S ¼ 1:17202  @ P A 
mm,i DT
i

5:321  103
D¼ þ 2:3376  102  9:297  101  T þ 1:417  103  T 2
T
 8:292  107  T 3
pffiffiffiffiffi
T DH ¼ 1 þ 1:5  I m
156 3 Seawater: Composition and Properties

P
mmi
i
I0 ¼
2
3:48662  102
B¼ þ 6:72817  9:71307  101  ln ðT Þ
T
40:5016
C¼  0:721404 þ 0:103915  ln ðT Þ
T
ρW ðT Þ ¼ 1:00157  1:56096  104  t  2:69491  106  t 2

The equation systems of the TEOS-10 standard also include calculation functions
for determining the osmotic coefficient and osmotic pressure (see Sect. 3.2,
Table 3.5).
A comparison of the osmotic coefficients at 25  C and for different salinities as
calculated with the above-presented algorithms (Fig. 3.33) reveals the following
picture:

• The data calculated with TEOS-10 and PHREEQC-Pister show very good agree-
ment for the entire range of salinity from 20 to 80 g/kg.
• Although the values obtained with the Bromley polynomial, in the range of
20–50 g/kg salinity, are slightly higher than the TEOS-10/PHREEQC-Pister
data, the difference is negligible for practical calculation of the osmotic pressure,
as it is a maximum of 0.2 units in the third decimal place in the range of the
osmotic coefficient from 0.90 to 0.92. In the range of >50–100 g/kg salinity, the
values agree with the TEOS-10/PHREEQC-Pister data.
• The values of the PERMASEP equation show good agreement with the TEOS-
10/PHREEQC-Pister data in the range of 20–40 g/kg salinity. Beyond that, the
calculation data differ significantly from those of the other equations and tend
toward lower values.

Of the compared modes of calculation, in terms of both simplicity of use and


agreement with the other calculation methods, the Bromley polynomial Eq. (3.116a)
and the function of the TEOS-10 standard can well be used to calculate the osmotic
coefficient for different seawater salinities and temperatures and also to calculate the
osmotic pressure.
In the graph presented in Fig. 3.34, the dependence of the osmotic coefficient on
seawater salinity and temperature was calculated with the Bromley polynomial. The
data belonging to the graph are given in Annex 3.A2, Table 3.30.
It can be seen from Fig. 3.34 that, in the relevant salinity range for membrane
desalination from 20 g/kg to a maximum of 80 g/kg and a temperature range from
10  C to a maximum of 50  C, the deviation of seawater from the behaviour of an
ideal solution is appreciable. Here, the osmotic coefficient has a range of
0.898–0.947. Consequently, the osmotic pressure obtained with Eq. (3.115) is
3.2 Properties of Seawater 157

significantly lower than the values calculated with the van‘t Hoff equation for ideal
solutions, i.e. without consideration of the osmotic coefficient.
The osmotic pressure of seawater for a salinity from 20 to 110 g/kg and a
temperature range of 10–70  C is presented in Fig. 3.35. The values in the graph
were calculated using Eq. (3.114), i.e. with consideration of correction factor ( fvw)
for the molar volume of water in seawater. This yields good agreement between the
thus obtained osmotic pressure and the values obtained with the function for osmotic
pressure according to the TEOS-1 standard. Without correction factor ( fvw), i.e. as in
Eq. (3.115), the TEOS-1 data for osmotic pressure are slightly lower in the first
decimal place.
Strictly speaking, however, the osmotic pressures presented in Fig. 3.35 are valid
only for seawater or seawater concentrates whose composition conforms to the
principle of constant proportions (see Sect. 3.1.1). Especially for concentrates from
membrane desalination processes, the composition of which tends toward an enrich-
ment of divalent ions owing to the stronger retention of those components by the
membranes, the osmotic coefficient should be calculated by means of the Pitzer
equations. This algorithm takes better account of the individual changes in compo-
sition of such concentrates.

Fig. 3.33 Osmotic coefficient—comparisons of different modes of calculation


158 3 Seawater: Composition and Properties

Fig. 3.34 Osmotic coefficient—dependence on salinity and temperature

Fig. 3.35 Osmotic pressure—dependence on salinity and temperature


3.2 Properties of Seawater 159

3.2.4 Chemical Equilibria in Seawater

The salts in the sea form a complex buffer system whose behaviour, together with
that system’s interaction with the carbon dioxide in the atmosphere, determines the
acidity and alkalinity, i.e. pH value, of seawater. The equilibria in this system are
influenced especially by the salts of weak acids, such as mainly carbonic acid and
boric acid and, to a lesser extent, also phosphates and silicates. The pH value of
seawater is determined primarily by the inorganic carbon cycle, in which carbon
dioxide, irrespective of its low concentration in the atmosphere, is of key importance
for the character of seawater. This is because, unlike the other gases in the atmo-
sphere, CO2 reacts chemically with water to form carbonic acid salts with different
degrees of dissociation and solubilities.
Under certain conditions, however, it is additionally necessary to take into
consideration the influence of the boric acid/borate equilibrium (although this is
significantly smaller than that of the CO2/carbonate equilibrium) on determination of
the pH value of seawater or desalination concentrates.
The buffer systems of seawater, and the therewith associated (as a function of the
pH value) distribution of their components, are of interest not only for oceanography
and the study of marine life but also for the design and dimensioning of membrane
systems for seawater desalination. Depending on the charge numbers of the
components of the CO2/carbonate and boric acid/borate equilibria, these constituents
are retained to different degrees by reverse osmosis membranes. Therefore, knowl-
edge of the characteristics of each equilibrium and the possibilities of influencing it
by changing the pH value are therefore important information for the design of
SWRO systems.

3.2.4.1 CO2/Bicarbonate/Carbonate Equilibrium


Water reacts with carbon dioxide (CO2) primarily in the following manner:

K0
CO2ðgÞ þ H2 O , CO2ðaqÞ
K
CO2ðaqÞ þ H2 O , H2 CO3 ¼ CO2

Undissociated carbonic acid (H2CO3) cannot be analytically distinguished from


dissolved CO2. Its concentration is less than 1/1000 of the concentration of dissolved
carbon dioxide [CO2(aq)]. As dissolved carbon dioxide, therefore, the sum of CO2
and H2CO3 is denoted as CO2 and, in consequence, its concentration is referred to as
[CO2 ].
Other reactions between dissolved carbon dioxide and water take place according
to the following equilibrium equations:

K1
CO2 þ H2 O , Hþ þ HCO
3
160 3 Seawater: Composition and Properties

K2
HCO þ
3 , H þ CO3
2

A further factor is the intrinsic dissociation of water:

Kw
H2 O , Hþ þ OH

Also for boric acid, there is an equilibrium between undissociated acid and the
anion borate according to

Kb
BðOHÞ3 þ H2 O , BðOHÞ þ
4 þ ½H 

The equilibria and concentrations of the individual components in this buffer


system can then be represented and calculated by the following equilibrium
constants:

½Hþ   HCO
3
¼ K 1 ð3:119Þ
CO2

½Hþ   CO2
3
¼ K 2 ð3:120Þ
HCO3

½Hþ   ½OH  ¼ K w ð3:121Þ

BðOHÞ þ
4  ½H 
¼ K B ð3:122Þ
BðOHÞ3

K 1 ¼ first stoichiometric dissociation constant of carbonic acid [mol/kg]


K 2 ¼ second stoichiometric dissociation constant of carbonic acid [mol/kg]
K w ¼ stoichiometric ion product of water [mol2/kg2]
K B ¼ stoichiometric dissociation constant of boric acid [mol/kg]

For the CO2/carbonic acid equilibrium in seawater, the calculation of the stoi-
chiometric constants for the equilibrium reactions as in Eqs. (3.119, 3.120, and
3.121) as well as the equilibrium constant of the boric acid equilibrium as in
Eq. (3.122) and its dependence on seawater salt content and temperature were
summarized in “UNESCO technical papers in marine science 51” (1987) [60]
according to the best available knowledge at the time. However, further research,
especially into how seawater acidity is affected by global warming and the
associated increase in the carbon dioxide concentration in the atmosphere, has
suitably broadened the available knowledge about the CO2/carbonic acid equilib-
rium of seawater. The following functions for calculating the first and second
dissociation constants of carbonic acid Eqs. (3.123 and 3.123a) are based on studies
and publications by Millero et al. in 2006 and 2010 [61, 62].
3.2 Properties of Seawater 161

Bi
pK i ¼ pK 0i þ Ai þ þ Ci  ln ðT Þ ½mol=kgsw  ð3:123Þ
T
Ai ¼ a0  S0:5 þ a1  S þ a2  S2

Bi ¼ a3  S0:5 þ a4  S

C i ¼ a5  S0:5
b1
pK 0i ¼ b0 þ þ b2  ln ðT Þ ð3:123aÞ
T

Dissociation constants in seawater


pH scale ¼ pHtotal
pK 1 pK 2
a0 +13.4051 +21.5724
a1 +0.03185 +0.1212
a2 5.218105 3.714104
a3 531.095 798.292
a4 5.7789 18.951
a5 2.0663 3.403
pK 1 pK 2
b0 126.34048 90.18333
b1 +6320.813 +5143.692
b2 +19.568224 +14.613358

Range of validity of equation: S ¼ 0–50 g/kg; t ¼ 0–50  C


The dependence of the respective negative logarithms pK1 and pK2 of the
dissociation constants on the temperature and salinity of seawater is presented in
Fig. 3.36 for the first dissociation constant and in Fig. 3.37 for the second dissocia-
tion constant of carbonic acid. The corresponding values of the two dissociation
constants are contained in Annex 3.A2, Table 3.31.
The ion product of water for a given temperature and as a function of seawater
salinity is calculated as in Eq. (3.124) [62]. Its dependence on seawater temperature
and salinity is presented in Fig. 3.38, and the corresponding values of the ion product
are given in Annex 3.A2, Table 3.32.

   
A A
ln K w ¼ A1 þ 2 þ A3  ln ðT Þ þ A4 þ 5 þ A6  ln ðT Þ  S0:5
T T
þ A7  S mol2 =kgsw 2 ð3:124Þ

A1¼+148.9652 A5¼+118.67
A2¼13,847.26 A6¼+1.0495
A3¼23.6521 A7¼0.01615
A4¼5.977
162 3 Seawater: Composition and Properties

Fig. 3.36 CO2/carbonate equilibrium-stoichiometric dissociation constant pK1 vs. temperature and
salinity

Fig. 3.37 CO2/carbonate equilibrium-stoichiometric dissociation constant pK2 vs. temperature and
salinity
3.2 Properties of Seawater 163

Range of validity of equation: S ¼ 0–45 g/kgSW; t ¼ 0–45 C ¼ 273.15–318.15 K


The pK value of the ion product is calculated from ln(K w ) as in Eq. (3.125). The
pH value is determined from the H+ ion concentration [H+] as in Eq. (3.126), and the
pOH value is calculated from the ion product and the pH value as in Eq. (3.127):

1
pK w ¼ log 10 ½K w  ¼   ln ðK w Þ ¼ 0:43429  ln ðK w Þ ð3:125Þ
ln ð10Þ

pH ¼  log 10 ½Hþ  ð3:126Þ

pOH ¼ pK w  pH ð3:127Þ

The equations for the dissociation constants of carbonic acid Eq. (3.123) and the
ion product of water Eq. (3.124) were correlated from test results with natural and
synthetic seawater to a maximum salinity of 40–50 g/kg (see also the information on
the validity of the respective equations). As membrane desalination of seawater can
produce concentrates with a salinity up to 80 g/kg, the values in the graphs and tables
were extrapolated to that salinity. The data above a salinity of 50 g/kg should,
therefore, be regarded only as a guide to how salinity can influence these physico-
chemical properties of seawater. For more accurate calculations in the range of
desalination concentrates, it is advisable to perform such calculations with the
thermodynamic constants of the respective equilibria together with the activity

pK*w [mol2/kgSW2]

14.0

13.5

13.0

12.5 20
30
40
50
60
70
80
12.0
5 15 25 35 45 55
Temperature [°C]

Fig. 3.38 Ion product of water—dependence on temperature and salinity


164 3 Seawater: Composition and Properties

coefficients of the seawater components. If a full analysis of the relevant seawater is


available, the activity coefficients can be determined with the Pitzer equations,
i.e. using the PHREEQC software tool and the pitzer.dat database. The same applies
to the following equations for the dissociation constant of boric acid and the
solubility product of calcium carbonate (see Sect. 3.2.3.2.1).
With knowledge of two sum parameters, namely, total alkalinity (AT) and total
dissolved inorganic carbon (CT) or DIC, boron concentration (BT), the dissociation
constants of carbonic acid and boric acid, and the ion product of water, it is possible
to calculate the concentrations of the main components of the seawater buffer
system. Total alkalinity is determined by acidimetric titration, while the CT value
is determined by acidifying a seawater sample and measuring the release of CO2 or,
alternatively, by acidimetric titration [63]. Total alkalinity is approximately equiva-
lent to methyl orange alkalinity (m-value) or acid capacity to pH 4.3 ¼ KS4.3 value of
freshwater analysis.
Total alkalinity is defined as follows:
X
AT ¼ HCO
3 þ 2  CO3
2
þ BðOHÞ 
4 þ ½OH  þ ½MAC
 ½H þ 
¼ mT ð3:128Þ

[C] ¼ molarity of component per mass of solution ¼ mol/kg of seawater [mol/kgsw]


¼ mMi,m
AT ¼ total alkalinity [mol/kg]
MAC ¼ minor anionic components [mol/kg]
mT ¼ total methyl orange alkalinity [mol/kg]

The so-called minor anionic components (MAC) make up a negligibly small


proportion of total alkalinity and can therefore be left out of consideration without
loss of accuracy when calculating the components and properties of the buffer
system:

A0T ¼ HCO
3 þ 2  CO3
2
þ BðOHÞ  þ 0
4 þ ½OH   ½H  ¼ mT ð3:128aÞ

A0T ¼ total alkalinity without MAC [mol/kg]


m0T ¼ methyl orange alkalinity without MAC [mol/kg]

The same applies to borate in the pH range of seawater from 5.5 to 8.0, because, in
the neutral range, only a small amount of boric acid dissociates into borate, for which
reason it accounts for only a small proportion of total alkalinity (see Figs. 3.42 and
3.43). Equation (3.128a) is then simplified to Eq. (3.128b):
3.2 Properties of Seawater 165

AT,C ¼ mT,C ¼ A0T  BðOHÞ


4

¼ HCO
3 þ 2  CO3
2
þ ½OH   ½Hþ  ð3:128bÞ

AT,C ¼ total carbonate alkalinity with [OH] and [H+] [mol/kg]


mT,C ¼ methyl orange carbonate alkalinity with [OH] and [H+] [mol/kg]

Such simplification, however, is possible without loss of accuracy only in the


above-mentioned neutral range of seawater. In the case of desalination concentrates,
in which boron has been enriched and the bicarbonate/carbonate content has often
also been reduced, such a simplification can lead to significant inaccuracies, for
example, in calculation of the pH value. In this case, it is better to keep the borate
content in the equation and to take out the H+ and OH ions. In the neutral pH range
of 6.0–9.0, both of these ions account for an even smaller proportion of total
alkalinity than borate:

K w
Δ ¼ ½OH   ½Hþ  ¼  ½H þ  ð3:128cÞ
½Hþ 

AT,B ¼ A0T  Δ ¼ HCO


3 þ 2  CO3
2
þ BðOHÞ
4 ¼ mT,B ð3:128dÞ

AT,B ¼ total carbonate/boron alkalinity [mol/kg]


mT,B ¼ methyl orange carbonate/boron alkalinity [mol/kg]

Leaving both the boric acid/borate equilibrium and the ion product of water out of
consideration, one obtains from Eq. (3.128b) the carbonate alkalinity AC according
to

AC ¼ AT,C  Δ ¼ HCO
3 þ 2  CO3
2
¼ mC ð3:129Þ

AC ¼ carbonate alkalinity [mol/kg]


mC ¼ methyl orange carbonate alkalinity [mol/kg]

This simplification with carbonate alkalinity (AC) is often used in the CO2/
carbonate equilibrium for engineering calculations. In this case, however, it must
be borne in mind that, outside of the pH range of 6–9 and in cases where the
contribution of borate to alkalinity is not negligible (see above), significant calcula-
tion inaccuracies can occur already starting from a pH value of a maximum of 8.0.
Therefore, it should always be made clear which alkalinity (AT, AT,C, AT,B, or AC) has
been used for calculations in the CO2/carbonate/borate seawater buffer system.
The total content of dissolved inorganic carbon CT (DIC) in the CO2/carbonate
equilibrium is calculated as follows:
166 3 Seawater: Composition and Properties

C T ¼ DIC ¼ CO2 þ HCO


3 þ CO3
2
ð3:130Þ

CT ¼ DIC ¼ total dissolved inorganic carbon content [mol/kg]

Alternatively, if the phenolphthalein alkalinity ( p-value, base capacity to


pH 8.2 ¼ KB8.2) is known, CT can be calculated from total carbonate alkalinity (AT,
C) as follows:

p ¼  CO2 þ CO2
3 þ ½OH   ½Hþ  ð3:131Þ

C T ¼ AT,C  p ¼ mT,C  p ð3:132Þ

p ¼ phenolphthalein alkalinity [mol/kg]

If the value of dissolved inorganic carbon CT is known, then, using the first and
second dissociation constants of carbonic acid, it is possible to calculate the distri-
bution of CO2, HCO3, and CO3 and their contributions to total carbon content CT in
the CO2/carbonate equilibrium as a function of the pH value and salinity of seawater
as follows:
!
K K  K
C T ¼ CO2  1 þ þ1 þ 1 þ 2 2 ð3:133Þ
½H  ½H 

K 1 K 1  K 2
Ψ1 ¼ 1 þ þ ð3:133aÞ
½Hþ  ½Hþ 
2

K1 2  K 1  K 2
Ψ2 ¼ þ þ ð3:133bÞ
½H  ½Hþ 
2

CO2 1
¼ F CO2 ¼ ð3:134Þ
CT Ψ1

HCO 1 K
3
¼ F HCO3 ¼  þ1 ð3:135Þ
CT Ψ1 ½H 

CO2 1 K  K
3
¼ F CO3 ¼  1 þ22 ð3:136Þ
CT Ψ1 ½H 

FCi ¼ distribution rate factor of CO2/carbonate equilibrium components

The calculated distribution rate factor FCi then indicates the contribution of each
component to the total carbon content in the CO2/carbonate equilibrium.
3.2 Properties of Seawater 167

Fig. 3.39 CO2/bicarbonate/carbonate equilibrium—distribution of equilibrium compounds vs. pH


and salinity

Fig. 3.40 CO2/bicarbonate/carbonate equilibrium—distribution of equilibrium compounds vs. pH


and temperature
168 3 Seawater: Composition and Properties

Table 3.14 Equations for CO2/carbonate equilibrium calculations [63]


Parameters
No. Available To be calculated Equation
1 CO2 +
, [H ] HCO2
3
½CO2 K 1
½Hþ 
2 CO2
3
½CO2 K 1 K 2
½Hþ 
2
 
3 CT CO2
K K  K 
 1 þ ½Hþ1  þ 1 þ 22 ¼ CO2  Ψ1
½H 
4 HCO2 +
3 , [H ] CO2 ½Hþ ½HCO2
3 
K 1
5 CO2 , HCO2 [H+] K 1 ½CO2 
3
½HCO2
3 

Ac, [H ] CO2
+
Ac ½Hþ 
2
6
K 1 ð½Hþ þ2K 2 Þ

7 HCO2 Ac ½Hþ 
3 ½Hþ þ2K 2
Ac K 2
8 CO2
3 ½Hþ þ2K 2
CT, [H+] CO2 C T ½Hþ 
2
9
½Hþ  þK 1 ½Hþ þK 1 K 2
2

10 HCO2 CT K 1 ½Hþ 
3
½Hþ  þK 1 ½Hþ þK 1 K 2
2

C T K 1 K 2
11 CO2
3
½Hþ  þK 1 ½Hþ þK 1 K 2
2

12 C T , CO2 , [H+] HCO2


3
½Hþ ðC T ½CO2 Þ
½Hþ þK 2
13 CO2 K 2 ðCT ½CO2 Þ
3
½Hþ þK 2
 
14 Ac K
2  C T  CO2  2 þ ½Hþ1 
15 Ac, CT, [H+] CO2 ½Hþ ð2C T Ac Þ
2½Hþ þK 1
16 Ac ð½Hþ þK 2 Þ
CT  ½Hþ þ2K 2
17 HCO2 K 1 ð2C T Ac Þ
3 2½Hþ þK 1
18 CO2
3 Ac  C T þ CO2
19 ½Hþ Ac þK 1 ðAc CT Þ
2½Hþ þK 1
20 Ac, CO2 , [H+] CO2
3
½Hþ Ac K 1 ½CO2 
2½Hþ 
h  i
21 CT 1 K
 Ac þ CO2  2 þ ½Hþ1 
2

22 Ac, CT, CO2 , HCO2


3 2  C T  Ac  2  CO2
23 CO2
3 Ac  C T þ CO2

Figures 3.39 and 3.40 present distribution rate curves for CO2, HCO3, and
CO32 and the dependence of their distribution rate factors (FCi) on pH value and
salinity as well as on pH value and temperature. The first graph is plotted for a
constant temperature of 25  C and the second graph for a constant salinity of 35 g/kg.
Figure 3.39 shows what significant influence the pH value has on the ratio between
the components of the CO2/carbonate equilibrium. As the pH value rises, the
3.2 Properties of Seawater 169

equilibrium increasingly shifts from the CO2/HCO3 side toward higher numbers of
CO32 ions. This shift is even more pronounced with increasing salinity in the
neutral and weakly alkaline range between pH 7.0 and 9.5. Yet also a temperature
rise promotes the transformation of bicarbonate ions into carbonate ions in this
range. Both have a significant influence on the solubility of calcium carbonate in
this pH range.
In addition, the three components of the CO2/carbonate equilibrium exhibit
different retention behaviours across reverse osmosis membranes. While almost all
the carbon dioxide passes the membrane, the CO32 ion is retained much better than
bicarbonate. As shown by the distribution rate curves in the two graphs, therefore,
total retention of dissolved inorganic carbon (CT) by desalination membranes
increases with rising pH value and increasing salinity and temperature.
Table 3.14 compiles a number of equations for CO2/carbonate equilibrium
calculations. They make it possible for unknown equilibrium components to be
calculated from two to three components whose concentrations are known. The
equations are derived from carbonate alkalinity (AC). They are valid, therefore, only
for the CO2/carbonate equilibrium and then only for the pH range in which the
influence of the H+ and OH– ions on alkalinity is negligible.

3.2.4.2 Boric Acid/Borate Equilibrium


The boric acid/borate equilibrium is a part of the seawater buffer system. The
concentration of the borate anion B(OH)4, which results from the dissociation of
boric acid, is also a component of the total alkalinity AT of seawater. The contribu-
tion of borate to total alkalinity is dependent on the pH value of the equilibrium,
i.e. on the degree of dissociation of boric acid:

BT ¼ BðOHÞ
4 þ BðOHÞ3 ð3:137Þ

BT  K B
AB ¼ BðOHÞ
4 ¼
  ð3:138Þ
½Hþ  þ K B

AB ¼ boron alkalinity [mol/kg]


BT ¼ total boron content [mol/kg]

The stoichiometric dissociation constant of boric acid K B and its dependence on


seawater temperature and salinity are calculated as follows [64]:
 
  B1 þ B2  S0:5 þ B3  S þ B4  S1:5 þ B5  S2
ln K B ¼
T
   
þ B6 þ B7  S þ B8  S þ B9 þ B10  S0:5 þ B11  S
0:5

 ln ðT Þ þ B12  S0:5  T ½mol=kgsw  ð3:139Þ


170 3 Seawater: Composition and Properties

B1¼8966.90 B7 ¼ +137.1942
B2¼2890. 53 B8 ¼ +1.62142
B3¼77.942 B9 ¼ 24.4344
B4¼+1.728 B10¼25.085
B5¼0.0996 B11 ¼ 0.2474
B6¼+148.0248 B12 ¼ 0.053105

Range of validity of equation: S ¼ 5–45 g/kgSW; t ¼ 0–45  C ¼ 273.15–318.15 K


Figure 3.41 shows the dependence of pKB on seawater temperature and salinity.
The associated values of stoichiometric dissociation coefficient (K B) are compiled in
Annex 3.A2, Table 3.33.
As in the case of the CO2/carbonate equilibrium, it is possible also for the
components of the boric acid/borate equilibrium to calculate a distribution rate factor
(FBi) Eqs. (3.140 and 3.141) showing the respective components of boric acid and
borate in the equilibrium as a function of seawater temperature and salinity:

BðOHÞ K
4
¼ F BðOHÞ4 ¼  þ B   ð3:140Þ
BT ½H  þ K B

BðOHÞ3 K
¼ F BðOHÞ3 ¼ 1   þ B   ð3:141Þ
BT ½H  þ K B

FBi ¼ distribution rate factor of boric acid/borate equilibrium components

Fig. 3.41 Boric acid/borate equilibrium-stoichiometric dissociation constant pKB vs. salinity and
temperature
3.2 Properties of Seawater 171

Fig. 3.42 Boric acid/borate equilibrium-distribution of equilibrium compounds vs. pH and salinity

Fig. 3.43 Boric acid/borate equilibrium-distribution of equilibrium compounds vs. pH and temperature
172 3 Seawater: Composition and Properties

Distribution rate curves for both components are presented as a function of pH


value and salinity in Fig. 3.42 and as a function of pH value and temperature in
Fig. 3.43.
Borate is retained considerably more effectively by desalination membranes than
undissociated boric acid. As demonstrated by the distribution rate curves, the borate
content in the equilibrium increases exponentially with increasing pH value starting
from a pH value of 8.0 and reaches its maximum at a pH value of around 10.2, at
which point the boron content of seawater is almost entirely in the form of borate.
Alkalization in a membrane desalination plant, therefore, can significantly improve
the retention of boron (BT) by reverse osmosis membranes.

3.2.4.3 pH in Seawater
Seawater in contact with the atmosphere at atmospheric pressure has a CO2 content
that is determined by the partial pressure of the gaseous carbon dioxide in the air.
The CO2 partial pressure of the atmosphere is currently 3.9  104 atm on average in
dry air and trending up. In this case, pCO2, the negative logarithm of partial pressure,
is 3.4089. To determine the carbon dioxide concentration in seawater CO2 , first of
all the solubility coefficient (hCO2 ) for the relevant temperature and salt content is
calculated as in Eq. (3.56). Thereafter, it is possible as in Eq. (3.49) with the partial
pressure of carbon dioxide at atmospheric conditions (PCO2 ) to determine the
saturation concentration of CO2 in seawater.
If the total alkalinity is additionally known, the pH value of the seawater can then
be calculated. Depending on whether, when calculating the pH value, one wishes to
include the contribution of the ion product of water, the borate concentration, or both
to total alkalinity, it will be necessary to take into consideration the value of AT,C, AT,
B, or A’T, respectively. Alternatively, the pH value can be calculated from the total
alkalinity together with the value of total inorganic carbon CT as in Eq. (3.142)
(in this case with AT,B, i.e. including the contribution of borate to total alkalinity):
! !
BT  K B ½Hþ  þ ½Hþ   K 1 þ K 1  K 2
2
CT ¼ AT,B   þ   ð3:142Þ
½H  þ K B ½Hþ   K 1 þ 2  K 1  K 2

If Eq. (3.142) is solved for the H+ concentration, one obtains a cubic equation in
the form of the following:
 
½Hþ    A þ ½Hþ   K 1  ðA  1Þ þ K B  ðA  BÞ þ ½Hþ 
3 2

 K 1  K B  ðA  B  1Þ þ K 1  K 2  ðA  2Þ þ K 1  K 2  K B
 ð A  B  2Þ
¼0 ð3:143Þ
3.2 Properties of Seawater 173

AT,B B
A¼ ;B ¼ T
CT CT
If in the pH calculation of seawater both the influence of borate and that of the
intrinsic dissociation of water are to be taken into account, this results in a system of
equations that includes the H+ concentration even to the fifth power.
The cubic equation including the contribution of borate to total alkalinity
Eq. (3.143) can be solved numerically by iteration using, for example, the
Newton-Raphson method. An alternative possibility is, after conversion to the
standard form of a cubic equation, to use Cardano’s method or other mathematical
rules for solving cubic equations [65].
If calculation of the pH is limited to the CO2/carbonate equilibrium, Eq. (3.142) is
simplified to the following:
!
½Hþ  þ ½Hþ   K 1 þ K 1  K 2
2
C T ¼ AC  ð3:144Þ
½Hþ   K 1 þ 2  K 1  K 2

Solving Eq. (3.144) for the H+ concentration leads to Eq. (3.145), which includes
+
H to only the second power. A quadratic equation is also obtained if, to determine
the H+ concentration, alkalinity (AC) and the carbon dioxide concentration [CO2]
are used:

AC  ½Hþ  þ ðAC  C T Þ  K 1  ½Hþ  þ ðAC  2  CT Þ  K 1  K 2


2

¼0 ð3:145Þ

AC  ½Hþ   K 1  CO2  ½Hþ   2  K 1  K 2  CO2 ¼ 0


2
ð3:146Þ

After the equations have been converted to the standard form of the quadratic
equation, it is then possible to calculate the H+ concentration and pH as follows:

½Hþ  þ p  ½Hþ  þ q ¼ 0
2
ð3:147Þ
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
þ p p 2
½H  ¼  þ q ð3:147aÞ
2 2
r ffi!
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

p p 2
pH ¼  log  þ q ð3:147bÞ
2 2
 
for AC , C T : p ¼ þK 1  1  CACT

 
2  CT
q ¼ þK 1  K 2  1 
AC
½CO2 
for AC , CO2 : p ¼ K 1  AC
174 3 Seawater: Composition and Properties

2  CO2
q ¼ K 1  K 2 
AC
If calculation of pH leaves out of consideration not only the contribution of borate
to total alkalinity and that of the H+ and OH ions of the intrinsic dissociation of
water but also the CO32 ion concentration, Eq. (3.148) is obtained (see also Eq. 3.5
in Table 3.14) and, therefrom, Eq. (3.149) results in the following:

CO2
½Hþ  ¼ K 1  ð3:148Þ
HCO 3

HCO
pH ¼ pK 1 þ log 3
ð3:149Þ
CO2

If Eq. (3.149) is used, however, this reduces the accuracy of calculation of the pH
value, particularly in the weakly alkaline range from a pH value of around 7.5, where,
especially in the case of high salinity, the carbonate ions can account for as much as
between 15% and 30% of total carbonate alkalinity (see Figs. 3.39 and 3.40). At a pH
value of 8.5, the contribution of CO32 ions to carbonate alkalinity then rises,
depending on salinity, from 25% to as high as 70%, in which case it should certainly
no longer be left out of consideration. Consequently, this simplified equation is
insufficiently accurate to calculate the pH of seawater or desalination concentrates in
the weakly alkaline to alkaline pH range.
However, neglecting the contribution of borate in the pH range above 8.0 can also
appreciably affect the accuracy of calculating the pH value. The average AC of
seawater is around 2.6  103 mol/kgSW. The average BT is around
4.3  104 mol/kgSW, of which, depending on temperature and salinity, between
20% and 35% is in the form of B(OH)4 at pH 8.0 and between 60% and 85% at
pH 9.0 (see Figs. 3.42 and 3.43). Whereas borate contributes only around 3% to total
alkalinity (AT,B) at pH 8.0, that value rises to as high as 12% at pH 9.0. Accordingly,
the accuracy of determining pH is affected if borate alkalinity is left out of consider-
ation in the higher alkaline range.
For engineering calculations of pH of seawater in the salinity range of 20–50 g/kg,
the accuracy of Eqs. (3.147 and 3.147b), i.e. with carbonate alkalinity and without
borate alkalinity, is sufficient. For seawater in this salt content range, the pH value is
normally between 7.5 and 8.5, the average being 8.1. Only rarely and then only
locally, usually in shallow water, pH values of up to 8.9 can be detected.
Figure 3.44 shows the pH of seawater and seawater concentrates with salinities in
the range of S ¼ 20–80 g/kg as a function of AC and at a seawater temperature of
20  C. The curves were calculated as in Eq. (3.147b) and for a range of carbonate
alkalinity from 1.5  103 to 4  103 mol/kg, as is the case for seawater. With
regard to the calculation data and curves for salinities above 50 g/kg, reference is
made to the remarks in connection with the equations for the first and second
dissociation constants of carbonic acid in Sect. 3.2.4.1.
3.2 Properties of Seawater 175

Fig. 3.44 pH of CO2 saturated seawater—dependence on salinity and carbonate alkalinity

3.2.4.4 Calcium Carbonate/Carbonate/Bicarbonate/CO2 Equilibrium


Similar to alkaline earth sulphates and calcium fluoride, calcium carbonate is a
low-solubility component of the seawater salt content. As for those compounds, a
thermodynamic solubility product K 0sp,CaCO3 can be derived from its solution
equation (see Sect. 3.2.3.2.1). Alternatively, a stoichiometric solubility product
(K sp,CaCO3 ) can be determined by solution experiments in natural or synthetic
seawater or can be calculated from the thermodynamic solubility product using an
appropriate activity coefficient (see Sect. 3.2.3.2.2).
However, calcium carbonate can be present in seawater in various modifications
with different crystal structures, i.e.:

• Calcite
• Aragonite
• Vaterite

This depends on the prevailing conditions when the calcium carbonate is


precipitated.
Sometimes, the less stable crystal structures aragonite and vaterite are
transformed after a certain time into the more stable calcite. The modifications also
have different solubilities and therefore also different thermodynamic and stoichio-
metric solubility products.
176 3 Seawater: Composition and Properties

The thermodynamic solubility products K 0sp,i of the three modifications of calcium


carbonate and their dependence on temperature are obtained as follows:
 
a
log K 0sp,i ¼ a1 þ a2  T þ 3 þ a4  log ðT Þ ð3:150Þ
T

CaCO3 modification
Coefficient Calcite Aragonite Vaterite
a1 171.90651 171.945 171.1295
a2 0.077993 0.077993 0.077996
a3 +2839.319 +2903.293 +3074.688
a4 +71.595 +71.595 +71.595

K 0sp,calc ¼ thermodynamic solubility product of calcite


K 0sp,arag ¼ thermodynamic solubility product of aragonite

Normally, engineering calculations of the solubility of calcium carbonate in


seawater assume the stable form calcite. Differences in results when the stoichio-
metric solubility coefficient is experimentally determined can, however, be
explained by the fact that aragonite or vaterite modifications, instead of calcite,
were measured under the prevailing experimental conditions.
According to the solubility equilibria

K spðaÞ
CaCO3 ðsÞ , Ca2þ þ CO2
3

K spðcÞ
CaCO3 ðsÞ , Ca2þ þ CO2
3

the calcium carbonate solubility is calculated from the stoichiometric solubility


product and the respective Ca2+ or CO23 concentration as follows:

Ca2þ  CO2
3 ¼ K sp,CaCO3,ðaÞ,ðcÞ ð3:151Þ

The stoichiometric solubility product itself as a function of the temperature and


salt content of the seawater is then obtained as follows [66]:
   
mol
log K sp ¼ log K 0sp,i þ Ai  S0:5 þ Bi  S þ C i  S1:5 ð3:152Þ
kgsw

178:34
Acalc ¼ 0:77712 þ 0:0028426  T þ
T
Bcalc ¼ 0:07711
3.2 Properties of Seawater 177

C calc ¼ þ0:0041249
88:135
Aarag ¼ 0:068393 þ 0:0017276  T þ
T
Barag ¼ 0:10018

C arag ¼ þ0:0059415

Range of validity of equation: S ¼ 5–44 g/kgSW; t ¼ 5–40  C ¼ 278.15–313.15 K


K sp,calc ¼ stoichiometric solubility product of calcite
K sp,arag ¼ stoichiometric solubility product of aragonite

Figure 3.45 shows the dependence of stoichiometric solubility product (K sp,calc )


(calculated with Eq. 3.152) on temperature and salinity for calcite. The
corresponding values of K sp,calc are contained in Annex 3.A2, Table 3.34.
As in the case of other low-solubility salts, with knowledge of the solubility
product of calcium carbonate and the respective concentrations of Ca2+ and CO32,
the calcium carbonate saturation of seawater can be calculated as shown in Sect.
3.2.3.2 as in Eqs. (3.88 and 3.89).

Fig. 3.45 Stoichiometric solubility product (pKsp) of calcium carbonate (calcite)—dependence on


salinity and temperature
178 3 Seawater: Composition and Properties

The influence of the carbonate concentration on the solubility of calcium carbon-


ate establishes a relationship between calcium carbonate and the CO2/carbonate
equilibrium. In Eq. (3.151), for example, the CO32 concentration can be substituted
by the concentration of bicarbonate (HCO3) Eq. (3.120). If the solubility reaction of
calcium carbonate is in equilibrium, i.e. neither solution nor precipitation is taking
place, then a so-called saturation pH (pHs ) can be defined as follows:

K 2
pHs ¼  log Ca2þ  log HCO
3  log  ð3:153Þ
K sp,CaCO3

pHs ¼ p Ca2þ þ p HCO  


3 þ pK 2  pK sp,CaCO3 ð3:153aÞ

pHs ¼ saturation (equilibrium) pH

From the saturation pH value (pHs ) and effective or in situ pH value (pHeff ) of the
seawater, it is then possible to determine the solution state of calcium carbonate by
calculating the saturation index as follows:

SI ¼ pHeff  pHs ð3:154Þ


 
SI ¼ pHeff  p Ca2þ þ p HCO3 þ pK 
2  pK 
sp,CaCO3 ð3:154aÞ

SI ¼ saturation index

If the SI is negative, CaCO3 is still in solution. On the other hand, a positive SI


indicates the possibility of calcium carbonate precipitation or solution
supersaturation.
Calculating the SI for standard seawater with a salt content of 35 g/kg, it becomes
apparent that, even without being concentrated in the course of the desalination
process, the seawater is already supersaturated with calcium carbonate (see
Fig. 3.46). Supersaturation increases with rising temperature.
The extent to which stoichiometric modelling can be used as a method for
assessing the scaling potential of seawater of high salinity or of desalination
concentrates is dependent on the degree to which reliable values are delivered by
the calculation values of the polynomials for the second dissociation constant of
carbonic acid and the solubility product of calcium carbonate for higher salinities
and temperatures. As previously mentioned, with regard to salinity, these equations
have been correlated mainly from experiments with seawater to a maximum salinity
of 40–50 g/kg. For higher salt concentrations, therefore, the reliability of these
equations is not guaranteed, in which case an alternative possibility is to calculate
the SI with the thermodynamic coefficients and activity coefficients of the solution
components. In this case, the saturation pH and the SI are calculated as in
3.2 Properties of Seawater 179

PH eff & PH s SI
8.2 1.00

8.1
pH eff Salinity = 35 [g/kg]
8.0 0.95

7.9

7.8 0.90

7.7

7.6 0.85

7.5
LSI
7.4 0.80
pH s
7.3

7.2 0.75

7.1

7.0 0.70

6.9

6.8 0.65
15 20 25 30 35 40 45 50
Temperature [°C]

Fig. 3.46 Equilibrium pH and SI of standard seawater—dependence on temperature

Eqs. (3.155, 3.155a, and 3.155b). The corresponding activity coefficients γi are
determined as shown in Sect. 3.2.3.2.1.

K 02
pHs ¼  log Ca2þ  log HCO
3  log  log γCa2þ
K sp,CaCO3
0

 log γHCO3  ð3:155Þ

pHs ¼ p Ca2þ þ p HCO


3 þ pK 2  pK sp,CaCO3 þ pγCa2þ þ pγHCO3 
0 0
ð3:155aÞ

SI ¼ pHeff
 
 p Ca2þ þ p HCO
3 þ pK 2  pK sp,CaCO3 þ pγCa2þ þ pγHCO3 
0 0

ð3:155bÞ

γCa2þ ¼ activity coefficient Ca2+


γHCO3  ¼ activity coefficient HCO3

Another possibility for determining the dissolution tendency of calcium carbon-


ate in seawater of higher salinity and in concentrates is to use of the Stiff-Davis
180 3 Seawater: Composition and Properties

Stability Index (S&DSI), which calculation is derived from the saturation index
equation. The equation also contains the calcium concentration [Ca2+] and the
carbonate alkalinity (AC). However, the solubility of calcium carbonate and its
dependence on temperature and salinity are taken into consideration not by the
solubility product of calcium carbonate and the second dissociation constant of
carbonic acid but by an empirical constant (KS&DSI) (see Eqs. 3.156, 3.156a, and
3.156b).

S&DSI ¼ pHeff  pHs ð3:156Þ

pHs ¼ p Ca2þ þ p½AC  þ K S&DSI ð3:156aÞ


 
S&DSI ¼ pHeff  p Ca2þ þ p½AC  þ K S&DSI ð3:156bÞ

S&DSI ¼ Stiff-Davis stability index


KS&DSI ¼ Stiff-Davis stability index constant

The S&DSI is a standard calculation method in RO seawater desalination, and its


application is defined in ASTM standard D4582, which proposes the use of this
calculation method starting from a salt content of 10,000 mg/l [67]. More details and
comments on how to apply stoichiometric modelling by usage of the SI and S&DSI
as well as thermodynamic modelling of CaCO3 scaling for determination of the
scaling potential of calcium carbonate during design of reverse osmosis desalination
systems can be found at Sects. 5.3.4.1, 5.3.4.1.3 and 5.3.4.1.4.
Annexes
Annexes

Annex 3.A1

Table 3.15 Seawater composition at different locations


Arabian sea Mediterranean sea Atlantic
Parameter Dimension Gulf of Oman Arabian gulf Red sea Eastern part Western part North Africa Singapore
min. max. min. max. min. max. min. max. min. max. min. max. min. max.

Temperature C 22 33 18 37 24 35 18 27 16 26 15 24 26 32
pH 7.9 8.2 8.0 8.3 8.1 8.3 7.5 8.1 7.8 8.5 7.8 8.2 7.8 8.4
Ca mg/l 395.0 505.0 440.0 580.0 451.0 560.0 391.0 450.0 441.0 455.0 286.0 457.0 715.0 811.0
Mg mg/l 1350.0 1583.0 1580.0 1800.0 1460.0 1620.0 1453.0 1580.0 1204.0 1427.0 922.0 1504.0 950.0 1160.0
Na mg/l 10,906.0 12,075.0 12,970.0 15,000.0 11,700.0 13,522.0 12,470.0 12,476.0 10,945.0 12,026.0 11,715.0 11,877.0 9900.0 10,834.0
K mg/l 395.0 420.0 400.0 600.0 425.0 560.0 450.0 463.0 336.0 410.0 307.0 407.0 625.0 800.0
Ba mg/l < 0.1 0.1 0.0 0.1 <0.02 0.4 0.0 0.0 < 0.01 < 0.01 <0.1
Sr mg/l 7.5 8.5 9.0 12.0 7.5 9.5 5.0 13.0 6.2 7.8 5.1 8.0
CO3 mg/l 1.9 8.5 1.5 15.3 2.2 10,8 0.6 4.5 1.3 14.9 0.9 8.6 1.1 9.4
HCO3 mg/l 146.0 161.0 100.0 200.0 104.0 151.0 128.0 195.0 149.0 163.0 110.0 195.0 101.5 119.0
SO4 mg/l 2860.0 3322.0 3200.0 3631.0 3000.0 3600.0 2810.0 3406.0 2840.0 3076.0 1643.0 3040.0 2070.0 2900.0
Cl mg/l 19,590.0 21,925.0 23,300.0 27,105.0 21,179.0 24,290.0 22,100.0 22,400.0 19,263.0 21,480.0 20,235.0 21,480.0 18,255.0 20,000.0
Br mg/l 65.0 75.0 78.0 84.8 72.0 78.0 63.4 66.5 65.0 69.0 69.0 75.5 65.0 75.0
F mg/l 0.6 2.2 1.0 2.5 1.0 1.5 0.7 0.8 1.0 1.6 1.0 1.6 0.6 2.0
NO3 mg/l 2.2 4.2 0.0 0.3 <0.1 0.2 < 0.5 < 0.5 0.3 4.1 <0.01 0.3
B mg/l 3.1 5.5 4.0 7.0 5.0 5.5 4.8 5.2 4.6 5.1 4.0 5.1 4.0 5.0
SiO2 mg/l < 0.7 0.2 0.7 0.2 0.5 0.2 0.4 < 0.2 0.4 0.4 2.3 0.2 4.7
TDS mg/l 35,722.3 40,095.0 42,083.8 49,038.7 38,406.9 44,398.6 39,871.7 41,047.4 35,254.9 39,141.0 35,299.8 39,065.0 32,692.5 36,728.4
181
182

Table 3.16 Seawater composition at different locations


Bay of bengal Indian ocean South Pacific Ocean Gulf of Mexico
Australia south east
Parameter Dimension South east India Australia west coast Australia south coast coast Australia east coast Chile north west coast South coast USA
min. max. min. max. min. max. min. max. min. max. min. max. min. max.

Temperature C 25 31 14 26 12 26 10 25 15 25 12 16 10 33
pH 7.5 8.5 8.1 8.3 8.0 8.2 7.4 8.9 7.5 8.7 7.5 8.0 7.9 8.3
Ca mg/l 361.0 450.0 400.0 490.0 446.0 476.0 373.0 498.0 406.0 510.0 441.0 528.0 336.0 480.0
Mg mg/l 1216.0 1275.0 1300.0 1500.0 1520.0 1530.0 1260.0 1920.0 1100.0 1500.0 1197.0 1452.0 1400.0 1400.0
Na mg/l 10,556.0 12,130.0 10,919.0 11,600.0 12,300.0 12,918.0 11,073.0 11,800.0 10,142.0 12,000.0 11,564.0 12,650.0 10,000.0 11,600.0
K mg/l 380.0 484.0 410.0 470.0 458.0 470.0 410.0 578.0 350.0 600.0 312.0 484.0 394.0 460.0
Ba mg/l < 0.01 0.0 <0.004 0.0 <0.005 0.0 6.0 8.0 0.0 0.0 0.0 0.0 0.0 4.0
Sr mg/l 3.7 10.2 6.7 8.4 8.0 9.5 5.2 8.0 6.6 9.7 13.0 21.0 2.2 9.9
CO3 mg/l 0.8 17.6 2.0 9.2 2.0 7.4 0.4 41.9 0.4 19.0 0.4 2.0 1.0 9.6
3

HCO3 mg/l 145.0 177.0 121.0 160.0 156.0 157.0 120.7 180.5 106.0 132.0 103.0 135.0 107.0 148.0
SO4 mg/l 2562.0 3072.0 2730.0 2880.0 3120.0 3270.0 2160.0 3550.0 2300.0 3584.0 2600.0 3009.0 2290.0 2543.0
Cl mg/l 18,805.0 21,240.0 19,600.0 21,294.0 22,181.0 23,107.0 20,100.0 22,400.0 18,051.0 21,536.0 20,384.0 22,788.0 18,669.3 21,237.1
Br mg/l 65.0 75.0 61.0 82.4 77.0 81.9 54.0 77.0 65.0 76.0 70.7 87.4 75.0 62.0
F mg/l 0.6 1.4 0.2 1.7 0.8 0.9 0.7 1.6 0.7 1.4 1.5 2.4 1.0 5.1
NO3 mg/l 0.6 1.6 <0.005 0.0 0.0 0.7 0.0 0.2
B mg/l 2.0 4.5 3.9 5.0 4.8 5.1 3.5 5.4 3.4 5.8 2.7 5.5 3.0 7.8
SiO2 mg/l 0.8 0.1 0.2 <1 <1 0.1 0.2 0.1 0.3 2.0 2.0 0.4 12.3
TDS mg/l 34,098.5 38,938.3 35,553.9 38,500.9 40,273.6 42,032.8 35,566.6 41,069.3 32,531.3 39,974.4 36,691.3 41,166.3 33,278.9 37,969.2
Seawater: Composition and Properties
Annex 3.A2
Annexes

Table 3.17 Specific heat capacity


Specific heat capacity [kJ/kg/K] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 4.197 4.136 4.078 4.022 3.968 3.916 3.866 3.818 3.772 3.728 3.686 3.646 3.608
20 4.189 4.133 4.078 4.025 3.973 3.923 3.875 3.829 3.784 3.741 3.700 3.660 3.622
30 4.184 4.130 4.078 4.027 3.978 3.930 3.883 3.838 3.794 3.752 3.711 3.671 3.633
40 4.181 4.130 4.079 4.030 3.982 3.935 3.890 3.845 3.802 3.760 3.719 3.680 3.641
50 4.181 4.131 4.081 4.033 3.986 3.940 3.895 3.851 3.808 3.767 3.726 3.686 3.647
60 4.183 4.133 4.085 4.038 3.991 3.945 3.901 3.857 3.814 3.772 3.731 3.691 3.651
70 4.187 4.138 4.090 4.043 3.996 3.951 3.906 3.862 3.819 3.777 3.735 3.695 3.655
80 4.194 4.145 4.097 4.049 4.003 3.957 3.912 3.868 3.824 3.782 3.740 3.699 3.658
90 4.203 4.154 4.105 4.058 4.011 3.964 3.919 3.874 3.830 3.787 3.745 3.703 3.663
100 4.215 4.165 4.116 4.068 4.020 3.973 3.927 3.882 3.838 3.794 3.751 3.710 3.668
110 4.229 4.179 4.129 4.080 4.031 3.984 3.937 3.892 3.847 3.803 3.760 3.718 3.676
120 4.246 4.194 4.144 4.094 4.045 3.997 3.950 3.904 3.859 3.814 3.771 3.729 3.687
183
184

Table 3.18 Specific enthalpy


Specific enthalpy [kJ/kg] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 42.11 41.59 41.03 40.41 39.73 38.97 38.15 37.25 36.28 35.24 34.16 33.04 31.91
20 84.00 82.83 81.65 80.45 79.21 77.93 76.60 75.22 73.79 72.32 70.82 69.30 67.78
30 125.83 124.07 122.33 120.60 118.85 117.09 115.29 113.47 111.62 109.74 107.85 105.96 104.07
40 167.62 165.33 163.07 160.85 158.63 156.41 154.18 151.94 149.69 147.42 145.15 142.89 140.64
50 209.41 206.61 203.87 201.18 198.51 195.86 193.21 190.57 187.92 185.27 182.62 179.98 177.36
3

60 251.20 247.92 244.72 241.57 238.47 235.39 232.33 229.28 226.23 223.19 220.14 217.11 214.11
70 293.04 289.28 285.62 282.02 278.48 274.97 271.49 268.02 264.55 261.08 257.63 254.18 250.74
80 334.93 330.70 326.57 322.51 318.52 314.56 310.63 306.70 302.79 298.87 294.96 291.05 287.15
90 376.91 372.18 367.57 363.03 358.55 354.11 349.69 345.28 340.87 336.46 332.04 327.61 323.18
100 419.00 413.75 408.61 403.55 398.55 393.58 388.63 383.68 378.73 373.75 368.76 363.75 358.72
110 461.22 455.41 449.70 444.07 438.49 432.94 427.40 421.84 416.27 410.66 405.02 399.34 393.63
120 503.60 497.17 490.84 484.57 478.35 472.14 465.93 459.69 453.42 447.10 440.72 434.28 427.78
Seawater: Composition and Properties
Annexes

Table 3.19 Thermal conductivity


Thermal conductivity [mW/mK] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 588.3 587.8 587.2 586.6 586.1 585.5 585.0 584.4 583.8 583.3 582.7 582.1 581.6
20 603.6 603.0 602.5 601.9 601.4 600.8 600.3 599.7 599.2 598.7 598.1 597.6 597.0
30 617.3 616.8 616.3 615.7 615.2 614.7 614.2 613.6 613.1 612.6 612.1 611.6 611.0
40 629.6 629.1 628.6 628.1 627.6 627.1 626.6 626.1 625.6 625.1 624.6 624.1 623.6
50 640.6 640.1 639.6 639.1 638.7 638.2 637.7 637.2 636.7 636.3 635.8 635.3 634.8
60 650.2 649.7 649.3 648.8 648.4 647.9 647.4 647.0 646.5 646.1 645.6 645.2 644.7
70 658.5 658.0 657.6 657.2 656.7 656.3 655.9 655.4 655.0 654.6 654.2 653.7 653.3
80 665.5 665.1 664.7 664.3 663.9 663.4 663.0 662.6 662.2 661.8 661.4 661.0 660.6
90 671.3 670.9 670.5 670.1 669.7 669.4 669.0 668.6 668.2 667.8 667.5 667.1 666.7
100 675.8 675.5 675.1 674.8 674.4 674.1 673.7 673.4 673.0 672.7 672.3 672.0 671.6
110 679.3 679.0 678.6 678.3 678.0 677.7 677.3 677.0 676.7 676.4 676.0 675.7 675.4
120 681.6 681.3 681.0 680.7 680.4 680.1 679.8 679.5 679.2 678.9 678.6 678.3 678.0
185
186

Table 3.20 Vapour pressure


Vapour pressure [bar] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.012 0.011
20 0.023 0.023 0.023 0.023 0.023 0.023 0.023 0.022 0.022 0.022 0.022 0.022 0.022
30 0.042 0.042 0.042 0.042 0.042 0.041 0.041 0.041 0.041 0.040 0.040 0.040 0.040
40 0.074 0.073 0.073 0.073 0.072 0.072 0.071 0.071 0.071 0.070 0.070 0.069 0.069
50 0.123 0.123 0.122 0.122 0.121 0.120 0.119 0.119 0.118 0.117 0.117 0.116 0.115
3

60 0.199 0.198 0.197 0.196 0.195 0.194 0.193 0.192 0.191 0.189 0.188 0.187 0.186
70 0.312 0.310 0.309 0.307 0.305 0.304 0.302 0.300 0.298 0.296 0.294 0.292 0.290
80 0.474 0.472 0.469 0.467 0.464 0.461 0.459 0.456 0.453 0.450 0.447 0.444 0.441
90 0.702 0.698 0.694 0.691 0.687 0.683 0.679 0.675 0.671 0.666 0.662 0.658 0.653
100 1.014 1.009 1.003 0.998 0.993 0.987 0.981 0.975 0.969 0.963 0.957 0.950 0.944
110 1.433 1.426 1.419 1.411 1.403 1.395 1.387 1.379 1.370 1.362 1.353 1.344 1.335
120 1.986 1.976 1.966 1.955 1.944 1.933 1.922 1.911 1.899 1.887 1.875 1.862 1.849
Seawater: Composition and Properties
Annexes

Table 3.21 Boiling point elevation


Boiling point elevation [K] at
Temperature Salinity [g/kg]

C 10 20 30 40 50 60 70 80 90 100 110 120
10 0.073 0.150 0.232 0.317 0.407 0.501 0.599 0.701 0.807 0.917 1.032 1.151
20 0.079 0.163 0.251 0.344 0.442 0.545 0.652 0.764 0.880 1.002 1.128 1.258
30 0.085 0.176 0.272 0.373 0.479 0.590 0.707 0.829 0.956 1.088 1.225 1.368
40 0.092 0.190 0.293 0.402 0.517 0.637 0.764 0.895 1.033 1.176 1.325 1.480
50 0.099 0.204 0.315 0.433 0.556 0.686 0.822 0.964 1.112 1.267 1.428 1.595
60 0.106 0.219 0.338 0.464 0.597 0.736 0.882 1.035 1.194 1.360 1.532 1.711
70 0.114 0.234 0.362 0.497 0.639 0.788 0.944 1.107 1.277 1.455 1.639 1.831
80 0.121 0.250 0.387 0.530 0.682 0.841 1.007 1.181 1.363 1.552 1.748 1.952
90 0.129 0.267 0.412 0.565 0.726 0.895 1.072 1.257 1.450 1.651 1.860 2.076
100 0.138 0.284 0.438 0.601 0.772 0.952 1.139 1.335 1.540 1.752 1.973 2.203
110 0.146 0.302 0.465 0.638 0.819 1.009 1.208 1.415 1.631 1.856 2.089 2.331
120 0.155 0.320 0.493 0.676 0.868 1.068 1.278 1.497 1.725 1.962 2.207 2.462
187
188

Table 3.22 Latent heat of evaporation


Latent heat of vaporization [kJ/kg] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 2477.2 2452.5 2427.7 2402.9 2378.1 2353.4 2328.6 2303.8 2279.0 2254.3 2229.5 2204.7 2180.0
20 2453.6 2429.0 2404.5 2379.9 2355.4 2330.9 2306.3 2281.8 2257.3 2232.7 2208.2 2183.7 2159.1
30 2429.8 2405.5 2381.2 2356.9 2332.6 2308.3 2284.0 2259.7 2235.4 2211.1 2186.8 2162.5 2138.2
40 2406.0 2381.9 2357.9 2333.8 2309.7 2285.7 2261.6 2237.6 2213.5 2189.4 2165.4 2141.3 2117.3
50 2382.0 2358.1 2334.3 2310.5 2286.7 2262.9 2239.0 2215.2 2191.4 2167.6 2143.8 2120.0 2096.1
3

60 2357.7 2334.1 2310.5 2287.0 2263.4 2239.8 2216.2 2192.7 2169.1 2145.5 2121.9 2098.3 2074.8
70 2333.1 2309.8 2286.4 2263.1 2239.8 2216.4 2193.1 2169.8 2146.4 2123.1 2099.8 2076.4 2053.1
80 2308.1 2285.0 2261.9 2238.8 2215.8 2192.7 2169.6 2146.5 2123.4 2100.4 2077.3 2054.2 2031.1
90 2282.6 2259.7 2236.9 2214.1 2191.3 2168.4 2145.6 2122.8 2100.0 2077.1 2054.3 2031.5 2008.7
100 2256.5 2233.9 2211.3 2188.8 2166.2 2143.7 2121.1 2098.5 2076.0 2053.4 2030.8 2008.3 1985.7
110 2229.7 2207.4 2185.1 2162.8 2140.5 2118.2 2095.9 2073.6 2051.3 2029.0 2006.7 1984.4 1962.1
120 2202.1 2180.1 2158.1 2136.1 2114.1 2092.0 2070.0 2048.0 2026.0 2003.9 1981.9 1959.9 1937.9
Seawater: Composition and Properties
Annexes

Table 3.23 Density


Density [kg/m3] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 999.5 1007.4 1015.2 1023.0 1030.9 1038.7 1046.5 1054.4 1062.2 1070.1 1077.9 1085.7 1093.6
20 998.0 1005.7 1013.4 1021.1 1028.8 1036.5 1044.1 1051.8 1059.5 1067.2 1074.9 1082.6 1090.2
30 995.6 1003.1 1010.7 1018.2 1025.8 1033.4 1040.9 1048.5 1056.1 1063.6 1071.2 1078.7 1086.3
40 992.2 999.7 1007.1 1014.6 1022.1 1029.5 1037.0 1044.5 1052.0 1059.4 1066.9 1074.4 1081.8
50 988.1 995.5 1002.9 1010.3 1017.7 1025.1 1032.5 1039.9 1047.3 1054.7 1062.1 1069.5 1076.9
60 983.2 990.6 998.0 1005.3 1012.7 1020.0 1027.4 1034.7 1042.1 1049.5 1056.8 1064.2 1071.5
70 977.8 985.1 992.5 999.8 1007.1 1014.5 1021.8 1029.1 1036.5 1043.8 1051.2 1058.5 1065.8
80 971.8 979.1 986.5 993.8 1001.1 1008.5 1015.8 1023.1 1030.5 1037.8 1045.1 1052.5 1059.8
90 965.3 972.6 980.0 987.3 994.7 1002.0 1009.4 1016.7 1024.1 1031.5 1038.8 1046.2 1053.5
100 958.3 965.7 973.1 980.5 987.8 995.2 1002.6 1010.0 1017.4 1024.8 1032.2 1039.6 1047.0
110 950.9 958.3 965.8 973.2 980.6 988.1 995.5 1003.0 1010.4 1017.8 1025.3 1032.7 1040.2
120 943.0 950.5 958.1 965.6 973.1 980.6 988.1 995.6 1003.1 1010.6 1018.1 1025.6 1033.1
189
190

Table 3.24 Dynamic viscosity


Dynamic viscosity [kg/m/s]  103 at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 1.306 1.330 1.355 1.382 1.412 1.443 1.476 1.511 1.548 1.586 1.627 1.670 1.714
20 1.002 1.021 1.043 1.065 1.089 1.114 1.140 1.168 1.197 1.227 1.259 1.292 1.326
30 0.797 0.814 0.832 0.851 0.871 0.891 0.913 0.936 0.960 0.984 1.010 1.037 1.064
40 0.653 0.667 0.683 0.699 0.716 0.734 0.752 0.771 0.791 0.812 0.833 0.855 0.878
50 0.547 0.560 0.573 0.587 0.602 0.617 0.633 0.649 0.666 0.684 0.702 0.721 0.740
3

60 0.466 0.478 0.490 0.502 0.515 0.528 0.542 0.556 0.571 0.586 0.602 0.618 0.635
70 0.404 0.414 0.425 0.436 0.447 0.459 0.471 0.484 0.497 0.510 0.524 0.538 0.553
80 0.354 0.364 0.373 0.383 0.393 0.404 0.415 0.426 0.437 0.449 0.462 0.474 0.487
90 0.315 0.323 0.331 0.340 0.349 0.359 0.369 0.379 0.389 0.400 0.411 0.422 0.434
100 0.282 0.289 0.297 0.305 0.313 0.322 0.331 0.340 0.350 0.359 0.369 0.380 0.390
110 0.255 0.262 0.269 0.276 0.283 0.291 0.299 0.308 0.316 0.325 0.334 0.344 0.354
120 0.232 0.238 0.245 0.251 0.258 0.265 0.273 0.280 0.288 0.297 0.305 0.314 0.323
Seawater: Composition and Properties
Annexes

Table 3.25 Kinematic viscosity


Kinematic viscosity [m2/s]  106 at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100 110 120
10 1.307 1.320 1.335 1.351 1.369 1.389 1.410 1.433 1.457 1.482 1.509 1.538 1.567
20 1.004 1.016 1.029 1.043 1.058 1.075 1.092 1.110 1.130 1.150 1.171 1.193 1.217
30 0.801 0.812 0.823 0.836 0.849 0.863 0.877 0.893 0.909 0.925 0.943 0.961 0.980
40 0.658 0.668 0.678 0.689 0.700 0.712 0.725 0.738 0.752 0.766 0.781 0.796 0.811
50 0.553 0.562 0.571 0.581 0.591 0.602 0.613 0.624 0.636 0.648 0.661 0.674 0.687
60 0.474 0.482 0.491 0.499 0.508 0.518 0.528 0.538 0.548 0.559 0.570 0.581 0.593
70 0.413 0.420 0.428 0.436 0.444 0.452 0.461 0.470 0.479 0.489 0.498 0.508 0.519
80 0.365 0.371 0.378 0.385 0.393 0.400 0.408 0.416 0.424 0.433 0.442 0.451 0.460
90 0.326 0.332 0.338 0.345 0.351 0.358 0.365 0.373 0.380 0.388 0.396 0.404 0.412
100 0.294 0.300 0.305 0.311 0.317 0.324 0.330 0.337 0.344 0.351 0.358 0.365 0.373
110 0.268 0.273 0.278 0.283 0.289 0.295 0.301 0.307 0.313 0.320 0.326 0.333 0.340
120 0.246 0.251 0.255 0.260 0.265 0.271 0.276 0.282 0.287 0.293 0.300 0.306 0.312
191
192

Table 3.26 Surface tension


Surface tension [mN/m] at
Temperature Salinity [g/kg]

C 0 10 20 30 40 50 60 70 80 90 100
10 74.22 74.47 74.66 74.82 74.95 75.07 75.17 75.26 75.35 75.42 75.49
3

20 72.74 73.03 73.25 73.44 73.59 73.73 73.85 73.96 74.05 74.14 74.22
30 71.19 71.53 71.78 71.99 72.17 72.32 72.46 72.58 72.69 72.79 72.88
40 69.60 69.96 70.25 70.49 70.68 70.85 71.01 71.14 71.26 71.38 71.48
Seawater: Composition and Properties
Annexes 193

Table 3.27 O2 solubility coefficient


O2 solubility coefficient hO2 ,v [mol/l.atm]  103
Temperature Salinity [g/kg]

C 0 10 20 30 40
10 1.7038 1.5989 1.5005 1.4081 1.3214
20 1.3862 1.3066 1.2315 1.1608 1.0941
30 1.1744 1.1114 1.0517 0.9953 0.9419
40 1.0311 0.9793 0.9301 0.8834 0.8390

Table 3.28 N2 solubility coefficient


N2 solubility coefficient hN2 ,v [mol/l.atm]  103 at
Temperature Salinity [g/kg]

C 0 10 20 30 40
10 0.8395 0.7835 0.7313 0.6826 0.6371
20 0.6959 0.6527 0.6122 0.5742 0.5385
30 0.6003 0.5654 0.5325 0.5015 0.4723
40 0.5364 0.5069 0.4790 0.4526 0.4277

Table 3.29 CO2 solubility CO2 solubility coefficient [mol/l.atm]  102 at


coefficient
Temperature Salinity [g/kg]

C 0 10 20 30 40
10 5.366 5.105 4.857 4.621 4.396
20 3.910 3.732 3.562 3.400 3.245
30 2.983 2.859 2.741 2.627 2.518
40 2.370 2.284 2.201 2.121 2.044
194

Table 3.30 Osmotic coefficient of seawater


Osmotic coefficient Ø
Salinity SA
Temperature g/kg

C 10 20 30 40 50 60 70 80 90 100
10 0.8977 0.8987 0.9011 0.9051 0.9106 0.9177 0.9262 0.9362 0.9478 0.9608
20 0.9004 0.9013 0.9039 0.9081 0.9139 0.9213 0.9303 0.9409 0.9531 0.9670
3

30 0.9021 0.9031 0.9057 0.9100 0.9161 0.9238 0.9332 0.9443 0.9571 0.9716
40 0.9030 0.9039 0.9065 0.9110 0.9172 0.9251 0.9349 0.9464 0.9597 0.9747
50 0.9029 0.9037 0.9064 0.9109 0.9172 0.9254 0.9354 0.9472 0.9609 0.9763
60 0.9020 0.9027 0.9054 0.9099 0.9163 0.9246 0.9347 0.9468 0.9607 0.9765
70 0.9002 0.9009 0.9035 0.9079 0.9144 0.9227 0.9329 0.9451 0.9592 0.9752
Seawater: Composition and Properties
Annexes

Table 3.31 CO2/carbonate equilibrium—stoichiometric dissociation constants K 1 and K 2


CO2/bicarbonate equilibrium in seawater - stoichiometric dissociation constants (pH scale ¼ pHTotal)
K 1 [106]
Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 0.8702 0.9746 1.0201 1.0639 1.1525 1.2502 1.3656 1.5075
20 293.15 1.1031 1.2378 1.2953 1.3499 1.4582 1.5756 1.7125 1.8795
25 298.15 1.2219 1.3752 1.4402 1.5017 1.6226 1.7522 1.9025 2.0848
30 303.15 1.3402 1.5146 1.5884 1.6579 1.7935 1.9375 2.1032 2.3031
40 313.15 1.5684 1.7934 1.8887 1.9781 2.1509 2.3312 2.5352 2.7783
50 323.15 1.7754 2.0623 2.1851 2.3003 2.5220 2.7505 3.0052 3.3047
K 2 [109]
Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 0.4284 0.5462 0.6147 0.6992 0.9543 1.4324 2.4088 4.5922
20 293.15 0.6263 0.7971 0.8934 1.0104 1.3587 2.0021 3.2961 6.1386
25 298.15 0.7505 0.9576 1.0728 1.2117 1.6220 2.3747 3.8784 7.1573
30 303.15 0.8943 1.1465 1.2850 1.4508 1.9367 2.8219 4.5794 8.3870
40 313.15 1.2502 1.6269 1.8301 2.0700 2.7615 4.0035 6.4433 11.6724
50 323.15 1.7142 2.2801 2.5824 2.9351 3.9358 5.7089 9.1605 16.4994
195
196

Table 3.32 Ion product of water


Ion product water
K W [1014 mol2/kgSW2]
Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 1.0818 1.3318 1.4436 1.5473 1.7322 1.8888 2.0189 2.1242
20 293.15 2.7869 3.5082 3.8396 4.1528 4.7266 5.2315 5.6692 6.0416
3

25 298.15 4.3322 5.5139 6.0639 6.5880 7.5598 8.4295 9.1972 9.8635


30 303.15 6.6025 8.4959 9.3880 10.2450 11.8521 13.3132 14.6241 15.7825
40 313.15 14.5110 19.0813 21.2855 23.4340 27.5500 31.3998 34.9562 38.1981
50 323.15 29.8115 40.0476 45.0926 50.0771 59.8147 69.1584 78.0143 86.3030
Seawater: Composition and Properties
Annexes

Table 3.33 Boric acid/borate equilibrium—stoichiometric dissociation coefficient of boric acid


Dissociation coefficient of boric acid
K B [109 mol/kgSW]
Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 1.3145 1.5573 1.6631 1.7582 1.9122 2.0110 2.0474 2.0178
20 293.15 1.7273 2.0610 2.2073 2.3392 2.5553 2.6979 2.7571 2.7281
25 298.15 1.9638 2.3547 2.5266 2.6821 2.9383 3.1101 3.1858 3.1595
30 303.15 2.2233 2.6815 2.8839 3.0676 3.3724 3.5801 3.6771 3.6561
40 313.15 2.8232 3.4562 3.7392 3.9983 4.4349 4.7438 4.9053 4.9081
50 323.15 3.5573 4.4416 4.8433 5.2152 5.8543 6.3256 6.5992 6.6567
197
198

Table 3.34 Stoichiometric solubility product of calcium carbonate


Stoichiometric solubility product of calcium carbonate (calcite) (K sp arag )
K sp arag 107 [mol2/kg2]
Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 2.2770 3.5989 4.3210 5.1063 6.9641 9.4273 12.8896 17.9829
20 293.15 2.2161 3.5593 4.3035 5.1190 7.0652 9.6679 13.3502 18.7983
25 298.15 2.1726 3.5220 4.2758 5.1053 7.0950 9.7694 13.5679 19.2070
30 303.15 2.1208 3.4728 4.2346 5.0768 7.1084 9.8542 13.7709 19.6075
40 313.15 1.9940 3.3387 4.1110 4.9735 7.0800 9.9630 14.1161 20.3583
50 323.15 1.8404 3.1596 3.9331 4.8069 6.9711 9.9759 14.3543 21.0014
Stoichiometric solubility product of calcium carbonate (aragonite) (K sp arag )
3

K sp arag 107 [mol2/kg2]


Temperature Salinity [g/kg]

C K 20 30 35 40 50 60 70 80
10 283.15 3.7895 5.7202 6.7881 7.9928 11.0898 15.7725 23.3766 36.4370
20 293.15 3.6127 5.5380 6.6162 7.8394 11.0021 15.8105 23.6570 37.2021
25 298.15 3.4972 5.4058 6.4818 7.7061 10.8821 15.7257 23.6513 37.3714
30 303.15 3.3658 5.2481 6.3167 7.5365 10.7116 15.5701 23.5436 37.3879
40 313.15 3.0633 4.8656 5.9038 7.0970 10.2265 15.0507 23.0196 36.9465
50 323.15 2.7231 4.4119 5.3998 6.5436 9.5690 14.2718 22.0971 35.8722
Seawater: Composition and Properties
Annex 3.A3
Annexes

Table 3.35 Pressure unit conversion table


Atmosphere
Pascal (Pa) Decibar (dbar) Bar (bar) Technical atmosphere (at) (atm) Torr (Torr) Pound per square inch (psi)
1 pa ¼1 N/m2 10–4 10–5 10.197  10–6 9.8692  10–6 7.5006  10–3 145.04  10–6
1 dbar 104 ¼ 105 dyn/cm2 0.1 0.10197 98.692  10–3 75.006 1.45037744
1 bar 100,000 10 ¼ 106 dyn/cm2 1.0197 0.98692 750.06 14.5037744
1 at 98,066.5 9.806 65 0.980665 ¼1 kg/cm2 0.967841 735.56 14.223
1 atm 101,325 10.1325 1.01325 1.0332 ¼ 1 atm 760 14.696
1 torr 133,322 1.3332  10–2 1.3332  10–3 1.3595  10–3 1.3158  10–3 ¼ 1 Torr 19.337  10–3
1 psi 6894.757894.757 0.68948 68.948  10–3 70.307  10–3 68.046  10–3 51.715 ¼1 lbf/in2
Example: 1 Pa ¼ 1 N/m2 ¼ 104 dbar ¼ 105 bar ¼ 10.197  106 at ¼ 9.8692  106 atm, etc.
199
200 3 Seawater: Composition and Properties

Table 3.36 Concentration unit conversion table


X
Y g/kg mol/kg g/kg H2O mol/kg H2O g/l mol/l
g/kg 1 M 1000S
1000 M  1000S
1000
1
ϱ
M
ϱ

M  1000
mol/ 1 1 1 1000S 1000S 1 1
M 1000 ϱM ϱ
kg
M
g/kg 1000
1000S M  1000S
1000 1 M 1
ϱ  1000S
1000
ϱ  1000S
1000

H2O
mol/ 1
M  1000S
1000 1000
1000S
1
M
1 1
ϱM  1000S
1000 1
ϱ  1000S
1000

kg
H2O
g/l ϱ ϱM ϱ  1000S
1000 ϱ  M  1000S
1000
1 M
ϱ ϱ
mol/ M
ϱ M  1000S
1000 ϱ  1000S
1000
1
M
1
l
M ¼ Mol mass of compound [g/mol]; ϱ ¼ solution density [kg/l]; S ¼ salinity SR or SA [g/kg],
Y ¼ XX conversion factor

References
1. Ludwig, H., Reverse Osmosis Seawater Desalination Volume 2, Springer, 2022
2. Millero, F.J., Feistel R., Wright, D.G., McDougall, T.J., “The Composition of Standard
Seawater and the definition of the Reference -Composition Salinity Scale,” Deep Sea Research,
vol. I, no. 55, pp. 50 - 72, 2008.
3. Unesco technical papers in marine science 36, “Tenth report of the joint panel on oceanographic
tables and standards,” 1981.
4. Unesco technical papers in marine science 37, “Background papers and supporting data on the
Practical Salinity Scale 1978,” 1981.
5. Unesco technical papers in marine science 44, “Algorithm for computation of fundamental
properties of seawater,” 1983.
6. McDougall T.J., Jackett D.R., Millero F.J., “An algorithm for estimating Absolute Salinity in
the global ocean,” Ocean Science Discussions, vol. 6, pp. 215 - 242, 2009.
7. American Society for Testing and Materials International ASTM, “ASTM D4195 - 08 Standard
Guide for Water Analysis for Reverse Osmosis and Nanofiltration Application,” 2008.
8. Boerlage S., “Measuring seawater and brine salinity in seawater reverse osmosis,” Desalination
& Water Reuse, pp. 26 - 30, November - December 2012.
9. Millero F.J, “History of the Equation of State of Seawater,” Oceanography, vol. 23, no. 3,
September 2010.
10. The International Association for the Properties of Water and Steam IAPWS, “Release on the
IAPWS Formulation 2008 for the Thermodynamic Properties of Seawater,” Berlin, 2008.
11. The International Association for the Properties of Water and Steam IAPWS, “Supplementary
Release on a Computationally Efficient Thermodynamic Formulation for Liquid Water for
Oceanographic Use,” Doorwerth, The Netherlands, 2009.
12. IOC, SCOR and IAPSO, Intergovernmental Oceanographic Commission, “The international
thermodynamic equation of seawater - 2010: Calculation and use of thermodynamic
properties,” in Manuals and Guidess No. 56, UNESCO, 2010.
13. McDougall T.J., Barker P.M, “Getting started with TEOS-10 and the Gibbs Seawater (GSW)
Oceanographic Toolbox,” SCOR/IAPSO WG127, 2011.
References 201

14. Feistel R., Wright D.G., Jackett D.R., Miyagawa K., Reissmann J.H., Wagner W., Overhoff U.,
Guder C., Feistel A., Marion G.M., “Numerical implementation and oceanographic application
of the thermodynamic potentials of liquid water, water vapour, ice, seawater and humidair - Part
1: Background and equations,” Ocean Science, vol. 6, pp. 633-677, 2010.
15. Feistel R., Wright D.G., Jackett D.R., Miyagawa K., Reissmann J.H., Wagner W., Overhoff U.,
Guder C., Feistel A., Marion G.M., “Numerical implementation and oceanographic application
of the thermodynamic potentials of liquid water, water vapour, ice, seawater and humidair - Part
2: The library routines,” Ocean Science, vol. 6, pp. 695 - 718, 2010.
16. Fichtner-Handbook, Hömig H.E, Seawater and Seawater Distillation, Vulkan - Verlag Essen,
1978.
17. Sharqawy M.H, Lienhard J.H, Zubair S.M, “Thermophysical properties of seawater: a review of
existing correlations and data,” Desalination and Water Treatment, vol. 16, pp. 354-380, 2010.
18. The International Association for the Properties of Water and Steam IAPWS, “Revised Release
on the IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water
Substance for General and Scientific Use,” Doorwerth, The Netherlands, 2009.
19. Wagner W., Pruß A., “The IAPWS Formulation 1995 for the Thermodynamic Properties of
Ordinary Water Substances for General and Scientific Use,” Journal of Physical and Chemical
Reference Data, vol. 31, no. 2, 2002.
20. Weiss R.F, “The solubility of nitrogen, oxygen and argon in water and seawater,” Deep-Sea
Research, vol. 17, pp. 721-735, 1970.
21. Benson B.B, Krause D., “The concentration and isotopic fractionation of oxygen dissolved in
freshwater and seawater in equilibrium with the atmosphere,” Limnology and Oceanography,
vol. 29(3), pp. 620 - 632, 1984.
22. Geng M., Duan Z., „Prediction of oxygen solubility in pure water and brines up to high
temperatures and pressures,“Geochimica et Cosmochimica Acta, Bd. 74, pp. 5631-5640, 2010.
23. Mao S., Duan Z., „A thermodynamic model for calculating nitrogen solubility, gas phase
composition and density of the N2-H2O-NaCl system,“ Fluid Phase Equilibria, Bd. 248, pp.
103 - 114, 2006.
24. Weiss R.F., “Carbondioxide in Water and Seawater: Solubility of a non-ideal gas,” Marine
Chemistry, vol. 2, pp. 203-215, 1974.
25. Duan Z., Sun R., “An improved model calculating CO2 solubility in pure water and aqueous
NaCl solutions from 273 to 533 K and from 2 to 2000 bar,” Chemical Geology, vol. 193, pp.
257 - 271, 2003.
26. Millero F.J., Schreiber D.R., “Use of the Ion Pair Model to estimate Activity Coefficients of the
Ionic Components of Natural Water,” American Journal of Science, vol. 282, pp. 1508 - 1540,
November 1982.
27. Garrels R.M, Thompson M.E., “A chemical model for seawater at 25 degrees C and one
atmosphere total pressure,” American Journal of Science, vol. 260, no. No. 1, pp. 57 - 66,
January 1962.
28. Kester D.R., Pytkowwicz R.M., “Theoretical Model for the Formation of Ion-pairs in Seawa-
ter,” Marine Chemistry, vol. 3, pp. 365 - 374, 1975.
29. Kester D.R., Pytkowicz R.M., “Sodium, Magnesium and Calcium sulfate Ion-pairs in Seawater
at 25C,” Limnology and Oceanography, vol. 14, no. 5, pp. 686 - 692, September 1969.
30. Johnson K.S., Pytkowicz R.M., “Ion Association of Cl- with H+, Na+, K+, Ca2+ and Mg2+ in
Aqueous Solutions at 25 C,” American Journal of Science, vol. 278, pp. 1428 - 1447,
December 1978.
31. Johnson K.S., Pytkowicz R.M., “Ion Association of Chloride and Sulphate with Sodium,
Potassium, Magnesium and Calcium in Seawater at 25 C,” Marine Chemistry, vol. 8, pp. 87
- 93, 1979.
32. Dickson A.G., Whitfield M., “An Ion- Association Model for Estimating Acidity Constants (at
25 C and 1 atm Total Pressure) in Electrolyte Mixtures related to Seawater,” Marine Chemistry,
vol. 10, pp. 315 - 333, 1981.
202 3 Seawater: Composition and Properties

33. Millero F.J., Hawke D.J., “Ion interactions of divalent metals in natural waters,” Marine
Chemistry, vol. 40, pp. 19 - 48, 1992.
34. Whitfield M., “A Chemical Model for the Major Electrolyte Component of Seawater Based on
the Brönsted-Guggenheim Hypothesis,” Marine Chemistry, vol. 1, pp. 251 - 266, 1973.
35. Whitfield M., “A comprehensive specific interaction model for seawater. Calculation of the
osmotic coefficient,” Deep-Sea Research, vol. 21, pp. 57 - 67, 1974.
36. Whitfield M., “An improved specific interaction model for seawater at 25 C and 1 atmosphere
total pressure,” Marine Chemistry, vol. 3, pp. 197 - 213, 1975.
37. Whitfield M., “The extension of chemical models for seawater to include trace components at
25 C and 1 atm pressure,” Geochimica et Cosmochimica Acta, vol. 39, pp. 1545 - 1557, 1975.
38. Crea F., Foti C., de Stefano C., Sammartano S., “SIT parameters for 1:2 electrolytes and
correlation with Pitzer coefficients,” Annali di Chimica, vol. 97, pp. 85 - 95, 2007.
39. Zemaitis J.F., Clark D.M., Rafal M., Scrivner N.C., Handbook of Aqueous Electrolyte Ther-
modynamics, Design Institute for Physical Property Data (DIPPR); American Institute of
Chemical Engineers (AIChE), 1986.
40. Bromley L.A., “Thermodynamic Properties of Strong Electrolytes in Aqueous Solutions,”
AIChE Journal, Vols. 19, No.2, pp. 313 - 320, 1973.
41. Bromley L.A., Singh D., Ray P., Sridhar S., Read s.M., “Thermodynamic Properties of Sea Salt
Solutions,” AIChE Journal, Vols. 20, No.2, pp. 326 - 335, 1974.
42. Pitzer K., “Thermodynamics of Electrolytes I. Theoretical Basis and General Equations,” The
Journal of Physical Chemistry, Vols. 77, No.2, pp. 268 - 277, 1973.
43. Pitzer K.S., Mayorga G., “Thermodynamics of Electrolytes II. Activity and Osmotic
Coefficients for Strong Electrolytes with One or Both Ions Univalent,” The Journal of Physical
Chemistry, Vols. 77, No.19, pp. 2300 - 2307, 1973.
44. Pitzer K. S., Mayorga G., “Thermodynamics of Electrolytes III. Activity and Osmotic
Coefficients for 2-2 Electrolytes,” Journal of Solution Chemistry, Vols. 3, No.7, pp. 539 -
546, 1974.
45. Pitzer K.S., Kim J.J, “Thermodynamics of Electrolytes IV. Activity and Osmotic Coefficients
for Mixed Electrolytes,” Journal of the American Chemical Society, Vols. 96, No.18, 1974.
46. Harvie C.E., Weare J.H., “The prediction of mineral solubilities in natural waters: the Na-K-
Mg-Ca-Cl-SO4-H2O system from zero to high concentration at 25  C,” Geochimica et
Cosmochimica Acta, vol. 44, pp. 981 - 997, 1980.
47. Harvie C.E., Möller N., Weare J.H., “The prediction of mineral solubilities in natural waters: the
Na-K-Mg-Ca-H-Cl-SO4-OH-HCO3-CO3-CO2-H2O system to high ionic strength at 25  C,”
Geochimica et Cosmochimica Acta, vol. 46, pp. 723 - 751, 1984.
48. Raju K., Atkinson G., “Thermodynamics of "Scale" Mineral Solubilities. 1. BaSO4(s) in H2O
and Aqueous NaCL,” Journal of Chemical and Engineering Data, Vols. 33, No.4, pp. 490 -
495, 1988.
49. Raju K.U.G, Atkinson G., “Thermodynamics of "Scale" Mineral Solubilities. 2.SrSO4(s) in
H2O and Aqueous NaCL,” Journal of Chemical and Engineering Data, vol. 34 No.3, pp. 361 -
364, 1989.
50. Raju K.U.G, Atkinson G., “Thermodynamics of "Scale" Mineral Solubilities. 3. Calcium
Sulfate in Aqueous NaCL,” Journal of Chemical and Engineering Data, vol. 35 No. 3, pp.
361 - 367, 1990.
51. Möller N., “The prediction of mineral solubilities in natural waters: A chemical equilibrium
model for the Na-Ca-Cl-SO2-H2O system, to high temperature and concentration,” Geochimica
and Cosmochimica Acta, vol. 52, pp. 821 - 837, 1988.
52. Spencer R.J, Möller N., Weare J.H., “The prediction of mineral solubilities in natural waters: A
chemical equilibrium model for the Na-K-Ca-Mg-Cl-SO4-H2O system at temperature below 25
C,” Geochimica and Cosmochimica Acta, vol. 54, pp. 575 - 590, 1990.
53. Nordstrom D.K., Jenne E.A., “Fluorite solubility in selected geothermal waters,” Geochimica
and Cosmochimica Acta, vol. 41, pp. 175 - 188, 1977.
References 203

54. U.S. Geological Survey USGS - Parkhurst D.L., Appelo C.A.J., “Users Guide to PHREEQC
(Version 2) (Equations on which the program is based) - Water Recources Investigation Report
99-4259,” 1999.
55. U.S. Geological Survey - Parkhurst D.L., Appelo C.A.J., “Description of Input and Examples
for PHREEQC Version 3-A Computer Program for Speciation, Batch-Reaction, One-Dimen-
sional Transport and Inverse Geochemical Calculations:U.S. Geological Survey Techniques
and Methods, book 6, chap. A43,” U.S. Geological Survey USGS, 2013.
56. Charlton S.R., Parkhurst D.L., “Modules Based on the Geochemical Model PHREEQC for Use
in Scripting and Programming Languages,” Computers and Geosciences, 2011.
57. Merkel B.J., Planer-Friedrich B., Groundwater Geochemistry - A Practical Guide to Modeling
of Natural and Contaminated Aquatic Systems, 2nd ed., Springer - Verlag, 2008.
58. Rushdi A.I., McManus J., Collier R.W., “Marine barite and celestite saturation in seawater,”
Marine Chemistry, vol. 69, pp. 19 - 31, 2000.
59. PERMASEP, “PERMASEP Products Engineering Manual,” in Bulletin 220 - Factors
influencing the capacity of RO devices, Wilmington, DuPont, 1992.
60. Unesco technical papers in marine science 51, “Thermodynamics of the carbon dioxide system
in seawater,” 1987.
61. Millero F.J., Graham T.B., Huang F., Bustos-Serrano H., Pierrot D., “Dissociation constants of
carbonic acid in seawater as a function of salinity and temperature,” Marine Chemistry, vol.
100, pp. 80 - 94, 2006.
62. Millero F.J., „Carbonate constants for estuarine waters,“ Marine and Freshwater Research, Bd.
61, pp. 139 - 142, 2010.
63. PICES Special Publication 3, IOCCP Report No.8, “Guide to Best Practices for Ocean CO2
Measurements,” 2007.
64. Dickson A.G., “Thermodynamics of the dissociation of boric acid in synthetic seawater from
273.15 to 318,15 K,” Deep-Sea Research, vol. 5, pp. 755-766, 1990.
65. Culberson C.H., “Calculation of the in situ pH of seawater,” Limnology and Oceanography,
vol. 25(1), pp. 150 - 152, 1980.
66. Mucci A., “The solubility of calcite and aragonite in seawater at various salinities, temperatures
and one atmosphere pressure,” American Journal of Science, vol. 283, pp. 789 - 799, 1983.
67. ASTM International, “D4582-10 Standard Practice for Calculation and Adjustment of the Stiff
and Davis Stability Index for Reverse Osmosis,” ASTM International, 2013.
Seawater Reverse Osmosis (SWRO) Plant:
General System Configuration, Basic Design 4
Parameters and Conditions, and Overall
Planning Process

4.1 SWRO General Systems Configuration

A seawater reverse osmosis (SWRO) plant generally consists of the core process,
i.e. membrane desalination and associated systems for supplying the desalination
plant with seawater of suitable quality, post-treating the product water according to
the quality requirements of the consumers, and ensuring that the concentrate pro-
duced during the desalination process is discharged back into the sea (see Fig. 4.1).
Other key systems of such a plant serve to store the product water and transfer it to
the consumer and to deliver, store, and prepare the necessary chemicals and auxiliary
substances and feed them to the processes. Another key part of an SWRO plant is its
energy supply, i.e. its connection to existing infrastructure or to dedicated power
supply systems.
Depending, in particular, on the type of seawater pretreatment systems as well as
site-specific environmental conditions and required environmental standards, it may
be necessary for the process wastewater from the plant to be treated before being
discharged into the sea, in which case the wastewater from membrane cleaning is
supplied to an appropriate wastewater treatment plant.
Seawater extraction can take place either directly from the sea by appropriate
extraction systems either close to the surface or from a greater sea depth in the form
of an open intake or indirectly by appropriate intake structures on or near the shore in
the form of subsurface extraction, in which case the seawater is prefiltered as it
passes through the layers of soil.
In the case of an open intake, the seawater is pretreated by suitable screening
systems before it reaches the desalination plant. This is not normally necessary in the
case of indirect or subsurface extraction, which also requires significantly fewer

Supplementary Information The online version contains supplementary material available at


https://doi.org/10.1007/978-3-030-81931-6_4.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 205
H. Ludwig, Reverse Osmosis Seawater Desalination Volume 1,
https://doi.org/10.1007/978-3-030-81931-6_4
206 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Seawater
extraction

Seawater screening
& pumping

Wastewater
treatment

Pretreatment Sludge
dewatering
Sludge to disposal
to outfall
Cleaning Cleaning effluent
RO process system treatment
Concentrate

Post-Treatment

Product water
storage & pumping
Plant outfall
Distribution
network

Fig. 4.1 Basic configuration of a seawater reverse osmosis (SWRO) plant

additional pretreatment measures before the seawater comes into contact with the
membranes (see also [23], Chap. 4 in Volume 2).
In the pretreatment system of the SWRO plant and its various treatment stages, the
seawater is treated to such an extent that the resulting water quality is suitable to allow
reliable operation of the reverse osmosis membrane systems of the desalination stage
(see also [23], Chap. 2 in Volume 2). Selection of the pretreatment processes depends
on the prevailing seawater quality at the site of the plant. Configuration of the
pretreatment system must take into consideration not only the values of the quality-
influencing parameters but also their seasonal and weather-related variations. This is
particularly the case if there is high biological activity in the seawater as well as
increased concentrations of particulate matter in case of bad weather events.
The reverse osmosis membrane desalination process consists primarily of a
seawater desalination stage with high-pressure membranes with high salt rejection (see
Sects. 5.1 and 5.1.2). Depending on seawater salt content, seawater temperature, mem-
brane type, and operating conditions of the desalination stage, it is possible to achieve
permeate salt concentrations in the range of around 200 mg/l to max. 500 mg/l TDS. The
4.1 SWRO General Systems Configuration 207

extent to which this product salt content with its composition of predominantly monova-
lent cations and anions is suitable for a given intended use of the desalinated water will
depend on the required quality standards of the consumer(s).
If the product water is to be used as boiler feed water in a power plant, for
industrial power and process-steam generation, or for industrial processes requiring
water with a lower salt content (pharmaceutical and electronics industries,
electroplating, etc.), it will be necessary to further reduce the salt concentration of
the permeate from a single-stage seawater reverse osmosis plant. The permeate from
the first seawater desalination stage can undergo post-desalination in a second RO
desalination stage with low-pressure brackish water membranes. This makes it
possible to achieve a final product salt content of between 20 and 100 mg/l TDS
depending on the salt content of the seawater desalination system, temperature, and
membrane type. Instead of post-desalination by RO membranes, however, a low-salt
product with a TDS of <1–5 mg/l can be obtained if the permeate from the seawater
desalination stage undergoes post-desalination in a post-treatment stage
(as presented in Fig. 4.1) using electrodialysis and/or ion exchange. Frequently,
such desalination processes are not employed until after the second stage of a
two-pass SWRO plant, i.e. in combination with post-desalination by RO membrane,
in order to obtain a product water of low salt content. In such a configuration, the
reverse osmosis process consists of a two-pass RO tract, while the post-treatment
stage presented in Fig. 4.1 consists of a non-membrane post-desalination stage.
There are, however, industries, such as the mining industry, for which the salt
content of the permeate from single-stage SWRO is sufficiently low for use in their
processes.
Instead of being used in industry or for power plants, however, reverse osmosis
seawater desalination finds much wider application for producing drinking water and
water for agricultural irrigation. For this reason, the further chapters of this book will focus
exclusively on the use of reverse osmosis seawater desalination for such applications.
To produce drinking water and water for agricultural irrigation, the post-treatment
stage consists of systems in which the composition of the product water from
desalination is adjusted to drinking water quality by the addition of chemicals
(lime, limestone, carbon dioxide, etc.). This usually increases the salt content of
the product water by around 100–200 mg/l (see [23], Chap. 3 in Volume 2). After
post-treatment of the product water, the drinking water itself should normally have a
salt content of between 200 and 500 mg/l TDS. This normally requires a two-pass
reverse osmosis seawater desalination process.
In the post-treatment stage or during storage and before transport to the consumer,
hygienization of the product water is accomplished by the addition of disinfectants
(chlorine, chlorine dioxide, or chloramine) or by physical disinfection processes.
However, it is not only the salt content (TDS) that determines the configuration of
the reverse osmosis desalination process when drinking water is produced from
seawater. In addition to the salt content, the quality of the drinking water from an
SWRO plant is usually dependent also on specified limit values for chloride,
bromide, or the boron concentration of the drinking water. Especially the required
boron concentration of the drinking water usually determines whether and to what
208 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

4 3 8
2

3 8a
4b 4b 2 2a
4a

1
5

5a
9
5b
9 6

7
10
7

Item Item
System System
No. No.
Seawater screening and Lime water production and
1 5a
pumping dosing
Carbon dioxide storage and
2 Open gravity filters 5b
dosing
2a Pretreatment dosing 6 Drinking water storage tank
Filtered water pumps Drinking water pumping
3 7
and cartridge filters station
Reverse osmosis hall
4 8 Wastewater treatment plant
(1st pass and 2nd pass)
4a Reverse osmosis dosing 8a Wastewater storage basin
Product water interme-
4b 9 Workshop and warehouse
diate tanks
Administration and control
5 Potabilization 10
room building

Fig. 4.2 Arrangement of processes and treatment systems at a large-scale SWRO plant (Courtesy:
Sydney Desalination Plant)

extent a second permeate post-desalination stage is required. This is because the


rejection of the RO membranes is significantly lower for boron compounds than for
other ionogenic substances in water (see Sects. 5.1.5 and 5.1.5.2.4).
The pretreatment stage filters out solids present in suspended or colloidal form in
the seawater. Such solids are then contained in enriched form in the wastewater from
the pretreatment stage or, depending on the type of pretreatment, also in the sludge
resulting from pretreatment. According to the quality of the SWRO feed water as
well as the type and operating mode of the pretreatment stage and also depending on
any required environmental standards for discharging the SWRO concentrate/waste-
water back into the sea, it may be necessary to treat the wastewater from the
desalination plant. Such a wastewater treatment stage filters out the solids contained
4.2 Basic Design Parameters and Conditions 209

in the wastewater from the pretreatment stage before the treated wastewater is
supplied to the outfall of the plant after being added to the concentrate from the
reverse osmosis process. The sludge produced in the wastewater treatment plant is
dewatered and sent to landfill (see also [23], Chap. 5, Sect. 5.2 in Volume 2).
Figure 4.2 shows the arrangement of the individual treatment processes and their
associated systems and infrastructure facilities within a large-scale SWRO plant with
a capacity of 250,000 m3/day.

4.2 Basic Design Parameters and Conditions

To design a seawater reverse osmosis (SWRO) plant, it is necessary to know a


number of basic parameters and to identify and determine various requirements for
the design and operating mode of the plant.
The most important basic parameters for plant design are:

• Seawater composition and its range of variation at the intended site of the plant
• Seawater quality and its dependence on seasonal and weather-related variations
• Seawater temperature and its variations at the intended extraction point for the
plant
• The required nominal production capacity of the plant (net output capacity)
• The required range of capacity of the plant and the timely availability of its
required minimum and maximum performance

Other important design parameters are:

• The product recovery rate of the reverse osmosis systems of the membrane
desalination process and the pretreatment stage of the SWRO plant
• The composition of the produced drinking water and its quality

The planning and design requirements with a significant influence on the systems
configuration of an SWRO plant include:

• The site of the plant and its connection to existing infrastructure and supply
systems, such as to an existing power plant or, in a hybrid system, to a dual-
purpose seawater desalination power plant, or whether the plant as a stand-alone
plant with feed-in to an existing drinking water supply system is extensively
autarkic.
• The operating mode and required flexibility of the production capacity,
i.e. whether the plant will be mainly in continuous operation for base load supply
or whether it will serve to meet peak load demand, with its output either following
the variations in drinking water consumption, being adjusted in line with the
available power supply (e.g. from renewable energy sources), or in line with
variations in power costs at different times of day.
210 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

• The power supply to the plant. What is of prime significance in this respect is
whether the power supply will be extensively uninterruptible as a no-break power
supply from an existing power grid or whether the plant will receive its power
from, for example, renewable energy sources (solar or wind) without backup from
conventional power generation facilities and, therefore, with highly variable
availability of power.
• The requirements of the future plant owner and/or operator with regard to the
maximum allowable energy consumption of the SWRO plant. Such requirements
have a significant influence on plant design and configuration, especially on the
design and configuration of the reverse osmosis systems.
• The storage of the produced drinking water and the integration of the SWRO
plant into the operating regime and supply system of the existing water supply
infrastructure.
• The applicable environmental standards with regard to disposal of the wastewater
and concentrate from the plant, noise reduction, and other requirements for
integration of the plant into the area surrounding the site.

4.2.1 Basic Design Parameters, Their Acquisition


and Determination

Table 4.1 presents the key basic design parameters for a reverse osmosis seawater
desalination plant. The individual parameters and parameter groups are ordered
according to the above-described categories. Additional data and indicators
supplementing the design parameter groups are given together with their units of
measurement.
The seawater-specific design parameters with regard to the composition and
quality of the seawater (Table 4.1, No. 1) must be appropriately determined at the
start of a membrane desalination project. This mainly includes an investigation
concerning existing data from, for example, oceanographic studies and data material
from environmental authorities. However, as such information is not specific to
membrane desalination, it will be necessary to conduct additional work programs to
take samples of seawater at the intended site of the plant, to analyse these samples
and check their quality, and to evaluate the analysis data. In addition to taking
samples, it will be necessary to make in situ measurements of seawater tempera-
ture (Table 4.1, No. 2). It is also advantageous to measure the conductivity and pH
value at the same time as the samples are taken.
The design parameters of the SWRO plant, such as its design capacity,
production capacity, and delivery capacity, the required production capacity range,
and required flexibility of operation of the plant (Table 4.1, No. 3), must be known at
the start of the project or must be determined in its early stages. The same applies to
the requirements with regard to the composition and quality of the drinking water
that the plant is to produce (Table 4.1, No. 5).
4.2 Basic Design Parameters and Conditions 211

Table 4.1 Basic design parameters of an SWRO plant


Basic design parameters
No. Category Parameter Unit
1 Seawater composition and quality range
1.1 Seawater composition range
1.1.1 Inorganic composition • Total cations and anions mg/l, mg/kg
• Total dissolved solids (TDS)
1.1.2 Scalants • Calcium Ca2+ mg/l, mg/kg,
• Strontium Sr2+ mmol/l,
• Barium Ba2+ mmol/kg
• Sulphate SO42
• Fluoride F
• Alkalinity
1.2 Seawater quality range
1.2.1 Inorganic foulants • Iron Fe mg/l, μg/l
• Manganese Mn
• Alumina Al
1.2.2 Particulate and colloidal Sum parameters and indicators for
fouling particulate and colloidal fouling:
• Suspended solids (SS) mg/l
• Turbidity NTU
• Silt density index (SDI) SDI units
• Modified fouling index (MFI) MFI units
• Particle size and distribution Mm
1.2.3 Organic fouling Sum parameters and indicators for
organic fouling:
• Dissolved organic carbon content mg C/l
TOC, DOC, POC cm1
• Ultraviolet absorbance UV254 l/mg,m
• Specific UV absorbance (SUVA) mg/l
• Hydrocarbons, oil and grease
1.2.4 Biological fouling Indicators for biological fouling:
• Nutrient status (nitrogen and mg/l, μg/l
phosphorus concentration, etc.) cfu/100 ml
• Bacterial count cfu/100 ml
• Faecal count μg/l
• Chlorophyll A cells/ml
• Algae count mg O2/l
• BOD5
2 Seawater temperature range

2.1 Temperature • Min. temperature C
• Max. temperature
• Temperature average
• Course of annual temperature
3 SWRO plant capacity
3.1 SWRO plant output, design, • Net output capacity CNO m3/d
and production capacity • Plant availability AP %
• Net design capacity CND m3/d
• Gross production capacity CGP m3/d
• Gross design capacity CGD m3/d
(continued)
212 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 4.1 (continued)


Basic design parameters
No. Category Parameter Unit
3.2 SWRO plant capacity range • Minimum net output capacity m3/d
CMinNO
• Maximum net output capacity
CMaxNO
4 Product recovery rate
4.1 Product recovery rate of • Product recovery rate of plant YT %
plant and systems • Product recovery rate of
pretreatment YPr
• Product recovery rate RO1 YRO1
• Product recovery rate RO2 YRO2
5 SWRO plant product composition and quality
5.1 Product composition • TDS mg/l
• Chloride Cl mg/l
• Bromide Br mg/l
• Boron B mg/l
• Calcium Ca2+ mg/l, ppm
• Alkalinity (HCO3, CO32) CaCO3
• pH mmol/l, mg/l,
• Chlorine ppm CaCO3
-
mg/l
5.2 Product quality • Turbidity NTU
• True colour TCU
• Taste and odour –
• Hygienic parameters
– Total coliforms cfu/100 ml
– E. coli cfu/100 ml
• Corrosion and calcium carbonate
precipitation parameters
– Langelier Saturation Index LSI units
(LSI)
– Calcium carbonate ppm CaCO3
precipitation potential (CCPP)
• Corrosion potential parameters
– (Cl + 2 SO4)/HCO3 mmol/mmol
– Ryznar Index Ryznar units

Process-specific parameters, such as the product recovery rate of the reverse


osmosis systems (Table 4.1, No. 4), must equally be established at the start of the
design process, and the bidder or plant supplier must make a preliminary selection of
suitable pretreatment processes based on knowledge of the existing seawater quality.
Selection of the pretreatment process will also provide knowledge of its probable
product recovery rate, which is then used as the basis for designing the SWRO plant.
Especially in the case of open intake extraction, selection of the pretreatment
process and its suitability for the entire range of seawater quality at the intended site
4.2 Basic Design Parameters and Conditions 213

should be verified and confirmed by operation of a pilot plant. Such pilot tests should
be conducted at the same time as the sampling and analysis programs, and both the
tests and the programs should be run over a sufficiently long period of time to ensure
that extreme values of water quality at the proposed site are measured both by
analysis and in the pilot tests.

4.2.1.1 Range of Composition and Quality of Seawater


The extent and details of knowledge of the seawater composition (Table 4.1, No.
1.1), i.e. the concentrations of the different inorganic components and their
variations over time at the seawater extraction site, are a key influencing factor for
the design quality of the reverse osmosis desalination processes of an SWRO plant.
The same applies to those parameters that are indicative of the seawater quality,
i.e. the inorganic, organic, and biological fouling potential of the seawater
(Table 4.1, No. 1.2). These quality parameters provide information on the type and
scale of required pretreatment of the seawater in order to achieve a suitable feed
quality to the membrane desalination process.

4.2.1.1.1 Seawater Data Assessment


As noted above, it is normally not possible to rely solely on existing data (oceano-
graphic measurements and environmental studies) in order to obtain the information
on seawater composition and seawater quality required for the design of an SWRO
plant. However, information from such sources, such as measurements of annual
variations in salinity and temperature as well as studies of microbial conditions,
measurements of concentrations of nutrients, of certain organic substances, and also
of heavy metals in the seawater, can serve to complement and confirm the results of
an additional sampling and analysis program.
The goal of such a sampling and analysis program must be to determine the range
of variation in the composition, temperature, and quality of the seawater at the
intended site of the SWRO plant or at the intended seawater extraction point and
to document this information for the subsequent planning process.
A precondition for achieving this objective is appropriate planning of the type and
frequency of seawater analyses at the intended site of the SWRO plant and by the
choice and type of sampling, i.e. sampling location and method of sampling.
According to the type of information to be obtained from the analysis data, a
distinction must be made between three groups with regard to sampling and analysis,
namely determination of:

• The seawater composition (Table 4.1, No. 1.1)


• The seawater quality with reference to particulate fouling (Table 4.1, Nos. 1.2.1
and 1.2.2)
• The seawater quality with regard to organic and biological fouling (Table 4.1,
Nos. 1.2.3 and 1.2.4)

Although the composition of seawater is subject to daily, monthly, and annual


variations, these are in most cases largely independent of seasonal influences. The
214 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

only exception to this is when the plant is sited in a river estuary, where there are
different freshwater outflows to the sea depending on the time of year. If the seawater
is extracted after bank filtration, the salt content of the feed to the SWRO plant can be
subject to significant variations as a result of freshwater inflow from onshore
groundwater strata.
A reverse osmosis seawater desalination plant must be capable of delivering its
capacity and product quality across the entire range of seawater composition and
seawater quality at the proposed site. Therefore, a seawater sampling and analysis
program must be planned and designed to measure not just the normal conditions but
also the minimum and maximum conditions at the seawater extraction point with
regard to the composition and quality of the seawater. As far as such a program is
concerned, this means that, especially in the case of direct seawater extraction and
challenging raw water conditions, the program can take several months, particularly
for the purpose of measuring weather-related and seasonal variations in quality.

Seawater Sampling
General guidelines for the planning and execution of seawater sampling are
contained in ISO standards ISO 5667-1 [1], ISO 5667-3 [2], and ISO 5667-9 [3].
Detailed descriptions of the procedure for seawater sampling from the surface and
deeper sea layers as well as the necessary techniques and equipment are given in
Grasshoff, Methods of Seawater Analysis [4].
In the case of direct extraction of seawater for operation of an SWRO plant, the
range of variation in its ionogenic composition and salt content is normally signifi-
cantly lower than the variation in seawater quality caused by seasonal and weather-
related influences.
Therefore, the sampling and analysis of seawater composition can be planned at
regular intervals. If the analysis results show that there is sufficient data on
fluctuations in seawater composition, then the frequency of sampling and analysis
can be appropriately adjusted or reduced.
To determine the particulate, organic, and biological quality of the seawater,
however, it is necessary to take account of weather-related and climatic influences.
Seawater contains increased amounts of solids and silt especially during autumn and
winter storms. The peaks in organic content and in biological activity are
concentrated above all on the summer months, during which the water reaches its
highest temperatures.
With regard to the significance of the analysis results, it is also of crucial
importance that the seawater samples should be taken from approximately the
same depth as the later seawater extraction point for the desalination plant. This
applies both to analysis of the quality parameters and to in situ measurement of
temperature (see also Chap. 3, Sect. 3.1.2), conductivity, pH value, and silt density
index (SDI). To determine those parameters whose values are significantly
influenced by marine conditions, the sampling protocol should make a record of,
for example, wind conditions, sea conditions, and tide conditions.
If, prior to its extraction, the seawater passes through onshore or seabed strata
(indirect extraction, subsurface extraction), then its quality variations in
4.2 Basic Design Parameters and Conditions 215

particulate, organic, and biological fouling will be greatly reduced by the physical
and biological filtration characteristics of those strata. Conversely, if the seawater
passes through soil, it can become enriched with substances that it did not originally
contain or that it contained only in small concentrations, such as increased
concentrations of heavy metals (iron, manganese, etc.) or silicic acid compounds.
Anaerobic biological degradation can result in sulphate being converted to sulphide.
Therefore, when planning to sample seawater from indirect extraction, it is
normally not necessary to give consideration to weather-related and seasonal
influences on the sampling sequence or main focus of sampling or, if so, only to a
minor extent. However, attention should be paid to the degree to which the quality of
the seawater is negatively affected by its passage through soil. Yet also in this case, it
should be ensured that the samples are taken from a source that is representative of
the later operation of the desalination plant, e.g. from a test borehole with the same
soil stratification and depth as the later seawater extraction point.
The reproducibility and quality of the analysis results of the seawater samples is
crucially influenced by the preservation of the samples immediately after they have
been taken and the appropriate adaptation of the preservation method to the parame-
ter(s) to be analysed. This applies in particular to the seawater quality parameters
and, more especially, to the indicators of organic and biological fouling (Table 4.1,
Nos. 1.2.3 and 1.2.4).

Seawater Analysis
Table 10 in the Annex 4.A1 to this chapter shows the structure and contents of a full
seawater analysis of the kind necessary for designing a reverse osmosis seawater
desalination plant.
To calculate the reverse osmosis membrane process of an SWRO plant and its
product quality and to design the product post-treatment and determine the compo-
sition of the produced process water or drinking water, it is necessary to conduct a
full analysis of the inorganic composition of the seawater as well as the relevant
physical parameters of temperature, pH value, and conductivity (see Annex 4.A1,
Table 10, No. 1). This involves measuring the entire range of concentrations of the
analysis parameters, i.e. minimum, maximum, and normal values, if possible for a
period of 1 year.
The same applies to the scope of analysis of those parameters that are character-
istic of the seawater quality (see No. 2 in Table 10, Annex 4.A1). These quality
parameters and their concentration ranges must be known in order to determine the
necessary pretreatment systems of the SWRO plant (see also [23], Chap. 2, Sects. 2.1
and 2.2 in Volume 2).
If the seawater sample is taken after the pretreatment stage of the SWRO plant,
then the chlorine content and the associated oxidation/reduction potential are impor-
tant analysis parameters. In order to protect the reverse osmosis membranes, it is
necessary that the maximum values of these parameters should not be exceeded (see
Sects. 5.1.3 and 5.1.4, Table 5.2).
Although the trace substances in No. 3 of this table normally have no effect on the
design and operation of an SWRO plant, they can provide important information on
216 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

the extent to which the seawater quality at the site of the plant can be influenced or
diminished by industrial effluent discharges or by pollution from shipping.
Guidelines and descriptions for carrying out the analyses required for the above-
mentioned parameters are contained both in Grasshoff, Methods of Seawater Analy-
sis [4], and, more recently, also in further collections of standard methods of water
and wastewater analysis [5–10] (see Table 4.2).
However, methods for the analysis of low-salinity water, as mainly contained in
the collections of standard methods in [6–10], require appropriate modifications in
methodology or in the selection of physical or chemico-physical detectors before
they can be used to determine various substances in seawater and concentrates.

Table 4.2 Standards for the analysis of water and wastewater


Abbreviation
Institution Name of standard of standard References
– Grasshoff, Methods of Seawater Analysis – [4]
ASTM International Section 11—Water and Environmental ASTM [5]
Technology—Water, Annual Book of
ASTM Standards, Volumes 11.01 and
11.02
American Water Standard Methods of the Examination of APHA [6]
Works Association Water and Wastewater
American Public
Health Association
Water Environment
Federation
United States Drinking Water Analytical Methods USEPA [7]
Environmental
Protection Agency
International ISO/TC 147 Water quality ISO [8]
Organization for
Standardization
European CEN/TC 230 Water quality EN [9]
Committee for EN ISO
Standardization
DIN Deutsches German Standard Methods for the DIN [10]
Institut für Examination of Water, Wastewater and DIN EN
Normung e.V. Sludge DIN EN ISO
NA 119 Water
Practice Standards
Committee
NA 119-01-03AA
Water examination
4.2 Basic Design Parameters and Conditions 217

However, especially the Annual Book of ASTM Standards, Volume 11.02 [5],
contains test methods for determining the main inorganic components of seawater
that have been specially devised and tested for seawater analysis.1,2,3,4,5,6,7
As methods for seawater analysis are contained in various standard works and
publications, it is necessary for the laboratories involved in a sampling and analysis
program to make it clear in their analysis reports which analysis methods they have
used (see Annex 4.A1, Table 10). Particularly where more than one laboratory is
involved in a program, this is the only way of ensuring that the results can be
compared and evaluated.
When the seawater composition is being determined, to check the plausibility of
the results, the principle of constant proportions (see Sects. 3.1 and 3.1.1) can be
applied using Eqs. 3.3 and 3.4. The quality of the results can be additionally verified
on the basis of a comparison of electroneutrality (see Total Dissolved Solids (TDS)
in the same section) using Eqs. 3.19, 3.20, 3.21, and 3.22.
For determination of the seawater quality, the particulate matter part of Annex 4.A1,
Table 10, No. 2, 2.1 contains information on the concentration of particulate
matter in the seawater and its consistency, i.e. the content of fine- and coarse-
grained substances. This information can be used to establish what amount of solids
must be captured by the pretreatment stage of the SWRO plant and how efficiently
the existing solids can be captured according to their particulate distribution profile.8
A turbidity measurement can serve as a simple method to determine the solids
content if a reliable relationship between turbidity and solids content can be
established for the seawater samples.9 Otherwise, the solids content must be more
reliably determined by filtration of the seawater.10
Normally, the heavy metals iron and manganese as well as aluminium (see metal
foulants in Annex 4.A1, Table 10, No. 2, 2.2) occur in seawater only as trace
substances. In the case of subsurface extraction, however, they can be present in
higher concentrations, in which case additional measures must be taken to capture
them. Iron can also be introduced into the pretreated seawater by coagulant/floccu-
lant dosing during pretreatment, and its concentration must be suitably monitored.

1
ASTM D3352 Test Method for Strontium Ion in Brackish Waters, Seawater and Brine.
2
ASTM D3561 Test Method for Lithium, Potassium and Sodium Ions in Brackish Water, Seawater
and Brines by Atomic Absorption Spectrophotometry.
3
ASTM D3651 Test Method for Barium in Brackish Waters, Seawater and Brines.
4
ASTM D3868 Test Method for Fluoride Ions in Brackish Waters, Seawater and Brines.
5
ASTM D3875 Test Method for Alkalinity in Brackish Waters, Seawater and Brines.
6
ASTM D4130 Test Method for Sulfate Ion in Brackish Waters, Seawater and Brine.
7
ASTM D4458 Test Method for Chloride Ions in Brackish Waters, Seawater and Brine.
8
APHA 2560-Particle Counting and Size Distribution.
9
ASTM D6855 Determination of Turbidity Below 5 NTU in Static Mode; ASTM D7315 Determi-
nation of Turbidity above 1 NTU in Static Mode; ISO 7027 Waters Quality—Determination of
Turbidity.
10
ASTM D5907 Test Methods for Filterable Matter (Total Dissolved Solids) and Non Filterable
Matter (Total Suspended Solids) in Water.
218 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

The colloidal fouling potential of seawater can be summarily determined by the


silt density index (SDI)11 or modified fouling index (MFI) (Annex 4.A1, Table 10,
No. 2, 2.3). Both of these are membrane filter test methods in which a seawater
sample is filtered at constant pressure for a predetermined time through a membrane
filter system. This measures how the amount of filtrate is reduced during the test
period. This can then be used to calculate the respective parameter, SDI or MFI, the
value of which indicates the degree of particulate fouling potential of the seawater
(see also Sects. 5.3.3 and 5.3.3.1).
The organic content in seawater is made up of highly different organic substance
groups and compounds. A fundamental distinction must be drawn between natural
organic matter (NOM) and organic compounds, which enter the seawater especially
in coastal regions due to industrial and municipal wastewater discharges and
shipping. The presence of NOM in seawater in coastal areas can also be attributable
to the inflow of onshore freshwater containing degraded natural biomass. In the sea
itself, NOM occurs as a result of the degradation and metabolization of marine flora
and fauna. The NOM substance group is a highly heterogeneous mixture of different
organic compounds, including humic acids, lignins, tannins, and many other ali-
phatic and aromatic organic substances.
The concentration of organic matter in seawater, or the concentration of inorganic
carbon, can be determined by the sum parameters of dissolved organic carbon
(DOC) and total organic carbon (TOC) (Annex 4.A1, Table 10, No. 2, 2.4). DOC
is calculated from TOC according to Eq. 4.1.

DOC ¼ TOC  POC ð4:1Þ

DOC ¼ dissolved organic carbon [mg C/l]


TOC ¼ total organic carbon [mg C/l]
POC ¼ particulate organic carbon [mg C/l]

Particulate organic carbon (POC) is the organic matter contained in the particulate
matter of seawater. To characterize the nature of the organic matter, the ultraviolet
absorbance (UVA) is measured at specific wavelengths and the quotient of UVA at a
wavelength of 254 mm and DOC is determined as the specific ultraviolet absorbance
(SUVA)12 (see Eq. 4.2).

UVA254
SUVA ¼  100 ð4:2Þ
DOC

11
ASTM D4189 Test Method for Silt Density Index (SDI) of Water.
12
US EPA (USEPA)—Method 415.3—Determination of Total Organic Carbon and Specific UV
Absorbance at 254 nm in Source Waters and Drinking Water.
4.2 Basic Design Parameters and Conditions 219

SUVA ¼ specific ultraviolet absorbance [l/mg,m]


UVA254 ¼ ultraviolet absorbance at 254 nm [cm1]

At a wavelength of 254 nm of UV light, NOM, such as humic acid or lignin,


exhibits substance-specific peaks in UV absorbance. Therefore, a UV254 measure-
ment can serve not only to determine the concentration of organic matter but also to
obtain information on the extent to which organic matter is of natural provenance or
the result of industrial or municipal wastewater discharges. Specific UV absorbance
(SUVA) has established itself in water and wastewater analysis as an indicator of the
relationship between NOM and the total organic matter concentration and is com-
monly used to monitor the efficiency of surface water treatment plants.
Although the nature of the organic matter in seawater can without doubt be
analysed more thoroughly and in greater detail using more sophisticated analytical
methods, the thereby obtained information is of doubtful value for deciding on the
required pretreatment processes inasmuch as the degree of removal of organic matter
from a given raw water by pretreatment can be predicted not so much in terms of
specific substances as in terms of an overall reduction. Therefore, when choosing the
pretreatment processes for an SWRO project, it is advisable to adopt a pragmatic
approach that is based on the above-mentioned parameters and in which, as far as
possible, information is gained from seasonal variations of the parameters in combi-
nation with particular weather-related and seasonal events (see also [23], Chap. 2,
Sect. 2.2.2 in Volume 2).
The situation is different when it comes to the operation of a pilot plant or the
monitoring of an existing pretreatment process, in which case it makes perfect sense
to employ additional analysis methods for certain organic substances known to have
a significant influence on membrane fouling, so that the reduction of those
substances by the pretreatment process can be measured. This applies, for example,
to so-called transparent exopolymer particles (TEP) and also to the humic acid
components of NOM (see also [23], Chap. 2, Sect. 2.1 in Volume 2).
Industrial and municipal wastewater discharges as well as pollution from
shipping can substantially diminish the quality of seawater. Particularly oil and
grease as well as petroleum hydrocarbons13 (Annex 4.A1, Table 10, No. 2, 2.5)
are synthetic organic foulants that, if coming into contact with separation
membranes, are capable of significantly deteriorating the product flow of the
membranes and their rejection rates within quite a short time. Membrane
manufacturers require the almost total elimination of such foulants from the feed
to membrane desalination processes. The operation of a membrane desalination
plant can also be negatively affected by other organic substances that enter the sea
as a result of municipal or industrial wastewater discharges. The existence of such
discharges at the site of an SWRO plant is indicated by the presence of surface-active
substances (tensides) in the seawater. Depending on the analysis method used to

13
ASTM D3921 Test Method for Oil and Grease and Petroleum Hydrocarbons in Waters.
220 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

detect them, these compounds are described as MBAS (methylene blue active
substances), CTAS (cobalt thiocyanate active substances), or DBAS (disulphine
blue active substances).14,15,16 MBAS, CTAS, and DBAS are anionic,
non-ionogenic, and cationic tensides, respectively (Annex 4.A1, Table 10, No.
2, 2.5).
Biological fouling is a key disruptive factor in the operation of a membrane plant
fed with surface water. This applies both to the pretreatment systems and especially
also to the reverse osmosis membrane process for seawater desalination. To assess
the biological fouling potential at the proposed site of an SWRO plant, it is necessary
to employ both the sum parameters according to Annex 4.A1, Table 10, No. 2, 2.5
and other indicators of substances in seawater that cause or promote biological
growth (Annex 4.A1, Table 10, No. 2, 2.6).
Eutrophic conditions, i.e. excessive supply of nutrients, are especially conducive
to biological growth, such as algal blooms and jellyfish blooms, in seawater.17 Such
events can place a considerable strain on the pretreatment stage of an SWRO plant
owing to a dramatic increase in the concentration of particulate matter in the feed to
the plant. Ammonia nitrogen18 as well as inorganic or organic nitrogen19 and
phosphorus20 are examples of such nutrients that promote biological growth, espe-
cially at higher water temperatures. The same applies to the concentration of natural
organic matter in seawater. This, too, acts as a nutrient for biological growth. The
biochemical oxygen demand (BOD)21 and the relationship between BOD and TOC
indicate how much of the existing organic matter is biodegradable, i.e. how much of
it can serve as a nutrient for biological growth.
The N-P nutrient content is also an indicator of possible municipal wastewater
discharges into the sea.
The other parameters contained in Annex 4.A1, Table 10, No. 2, 2.6

• Chlorophyll A
• Pheophytin
• Algae count

14
APHA 5540 Surfactants.
15
ISO 7875-1 Water quality—Determination of surfactants—Part 1: Determination of anionic
surfactants by measurement of the methylene blue index (MBAS); ISO 7875-2 Water quality—
Determination of surfactants—Part 2: Determination of non-ionic surfactants using Dragendorff
reagent.
16
DIN 38409-20 German standard methods for the examination of water, wastewater and sludge;
parameters characterizing effects and substances (group H); determination of substances that react
with disulfine blue (H 20).
17
Grasshoff, Methods of Seawater Analysis, Section 10—Determination of nutrients.
18
ASTM D1426 Standard Test Methods for Ammonia Nitrogen in Water.
19
ASTM D3590 Standard Test Methods for Total Kjeldahl Nitrogen in Water.
20
APHA 4500-P Phosphorus; APHA 4500-P J. Persulfate Method for Simultaneous Determination
of Total Nitrogen and Total Phosphorus.
21
APHA 5210 Biochemical Oxygen Demand.
4.2 Basic Design Parameters and Conditions 221

• Total bacteria count


• Faecal bacteria count

provide an indication of the degree of biological growth in seawater. Chlorophyll


A and pheophytin are indicators of the phytoplankton concentration in seawater,
i.e. its content of herbal microorganisms. Chlorophyll A is a magnesium-containing
pigment for photosynthesis and is contained not only in phytoplankton and algae but
also in cyanobacteria. Therefore, its measurement makes it possible to determine the
concentration of biomass in seawater. Pheophytin is the main degradation product of
chlorophyll A, and its measurement provides an indication of the amount of no
longer reactive phytoplankton. Both parameters are determined in seawater prefera-
bly by means of a fluorescence measurement.22 Together with determination of the
algae count by direct counting, it is possible to establish the current status of
biological growth in a seawater sample. By tracking the variations of these
parameters over time, it is then additionally possible to determine the biological
growth potential. A sharp increase in the three above-mentioned parameters as well
as in turbidity points, for example, to the beginning of an algal bloom.
The total bacteria count indicates the status of bacterial growth, which, together
with plant biological growth as well as the released natural organic matter as nutrient
and mucus-forming metabolites from both, is the cause of membrane biofouling. The
type of faecal bacteria and the faecal bacteria count provide information on the extent
to which bacterial growth in seawater is initiated by municipal wastewater
discharges.
Consequently, to be able to estimate the membrane biofouling potential of
seawater and, therefore, the type and scale of the required pretreatment processes,
it is necessary to take account of a number of seawater quality parameters, such as
the parameters for:

• Organic matter content (Annex 4.A1, Table 10, No. 2, 2.4)


• Organic pollution from municipal and industrial discharges into the sea (Annex 4.
A1, Table 10, No. 2, 2.5)
• The biological fouling potential of the seawater from vegetable and bacterial
organic growth (Annex 4.A1, Table 10, No. 2, 2.6)

Documentation of the sampling conditions in a sampling and analysis program


should allow a correlation of the analysis parameters with seasonal and weather-
related conditions, so that the variations in the analysis data can be set in relation to
marine and weather-related events.
With regard to selection of the pretreatment processes according to the data on the
quality and fouling potential of seawater, reference is made to [23], (Chap. 2, Sect.
2.2.2.1 in Volume 2).

22
APHA 10200-H Chlorophyll; USEPA Method 445.0 In Vitro Determination of Chlorophyll a
and Pheophytin a in Marine and Freshwater Algae by Fluorescence.
222 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

In addition to the main inorganic substances and the parameters affecting the
seawater quality at the particulate, organic, and biological level, seawater also
contains a multiplicity of different trace substances, such as heavy metals and
organic compounds. These substances, which are usually present only in the micro-
gram range, normally have no influence on the design and operation of an SWRO
plant. However, the spectrum of trace substances in seawater can also include
substances that are classified in WHO guidelines and in national drinking water
standards as being of health significance and which are subject to compulsory
monitoring in the case of drinking water. The inorganic substances in question are
various heavy metals (see Annex 4.A1, Table 10, No. 3, 3.1)23 as well as the anion
cyanide (CN). Although heavy metals are usually present in seawater in
concentrations of just a few micrograms per litre, they can occur in higher
concentrations:

• In case of direct extraction as a result of coastal wastewater discharges into the sea
• In case of indirect (subsurface) extraction as a result of passage through soil strata
or inflow of contaminated groundwater

The same applies to organic substances, such as polycyclic aromatic


hydrocarbons (PAH) and trihalomethanes (THM), pesticides, and other compounds
that are subject to compulsory monitoring and which, as far as drinking water is
concerned, can be described as health-influencing substances.
Normally, reverse osmosis seawater membranes are highly efficient at capturing
the above-mentioned heavy metals as well as organic matter, and there is hardly any
risk of them entering the product water from an SWRO plant.
Nevertheless, those substance groups that are subject to compulsory monitoring
under drinking water standards should be included in a sampling and analysis
program and should be monitored over a certain period of time. If it turns out that
the seawater contains concentrations that are significantly above the locally normal
trace level or considerably exceed the monitoring values for drinking water, then, in
the case of an SWRO plant for drinking water, appropriate significance must be
attached to the monitoring of such substances in connection with the quality control
of the produced drinking water.

4.2.1.1.2 Pilot Plant Testing


Based on the analysis data for the particulate, organic, and biological seawater
quality parameters, it is possible to select pretreatment processes that can be
expected to produce the feed water quality required by the reverse osmosis
membranes of the desalination process (see [23], Chap. 2, Sect. 2.2.2.1 in Volume
2). For the selected processes or chains of processes, it is then necessary to establish
more detailed design, operation, and consumption parameters. In addition, it is

23
Grasshoff, Methods of Seawater Analysis, Section 12—Determination of trace elements; APHA
3112 Metals by Cold-Vapor Atomic Absorption Spectrometry; APHA3113 Metals by Electrother-
mal Atomic Absorption Spectrometry.
4.2 Basic Design Parameters and Conditions 223

necessary to examine whether the selected process with its projected parameters is
capable of achieving the feed water quality required for operation of the reverse
osmosis processes of the SWRO plant across the entire range of seawater quality.
If the treatment processes need to include coagulation/flocculation stages, it is
necessary to determine the consumption and design parameters, e.g. consumption of
coagulants and flocculants, as well as the process-specific design parameters for the
mixing and coagulation/flocculation conditions.
A design parameter for sedimentation and flotation processes is the surface
loading rate, while, for granular media filtration processes, it is the filtration rate.
For membrane filtration, the design parameters are membrane flux and transmem-
brane pressure.
An operating parameter for both granular media filtration and membrane filtration
is the length of the filter cycle between backwashing. This is dependent on the filter
rate and the solids load in the filter intake and therefore varies according to the
particulate quality of the seawater.
In the case of granular media filtration (dual media or multiple media filtration),
not only the length of the filter cycle but also the filtrate quality and the infiltration
duration are determined by the choice of filter materials and their grain size
distribution.
Further details on the design of the pretreatment processes as well as their design
and operating parameters can be found in [23], Chap. 2, Sect. 2.3 in Volume 2.
To be able to determine these design and operating parameters that are required
for the design of the pretreatment processes, either the design variables must be
known from experience of existing plants with similar seawater quality conditions or
the data must be determined by further investigation of the specific conditions at the
proposed site of the SWRO plant.
In bench-scale tests with seawater samples, comparative jar tests24 can serve to
determine guideline values for the chemical consumption of the coagulation/floccu-
lation stages. By extension of these tests, it is also possible to measure the sedimen-
tation and flotation rates of the flocs. These data can then be used to estimate the
possible surface load of the separation processes.
However, the information obtained from such bench-scale tests is not sufficient to
serve as a basis for the reliable design of the pretreatment processes. As such tests are
normally conducted with individual seawater samples, they can do no more than
give a snapshot of the seawater quality at the time the samples were taken. To obtain
results that are extensively representative of the performance and efficiency of a
pretreatment stage in the case of variations in seawater quality and that can also serve
as a basis for contractual assurances of product capacity, consumption, and quality, it
is necessary to conduct pilot plant tests. This is especially the case with seawater of
inferior and highly variable quality. For such conditions, pilot plant testing has
become established as an integral part of the plant design process, particularly for
large-capacity SWRO plants.

24
ASTM D2035 Standard Practice for Coagulation-Flocculation Jar Test of Water.
224 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Consequently, the purpose of such a pilot plant is:

• To compare possible pretreatment processes or chains of processes for their


suitability and efficiency
• To optimize the design and operating parameters of the selected pretreatment
process so that the required quality parameters for the feed to the RO membrane
desalination process (SDI, MFI, turbidity, and organic sum parameters) can be
achieved across the entire seawater quality range (for conventional filtration
processes, the choice of filter media as well as their grain size can also be part
of process optimization during pilot plant testing)
• To determine and optimize the amounts of chemicals required for operation at
different feed conditions
• To test the performance of the RO membranes when fed with the pretreated
seawater (variations in operating pressure, differential pressure, permeate output,
and permeate quality) and to determine the frequency and type of the chemical
cleaning operations required for the membranes

Normally, the design and operating parameters of the desalination process need
not be determined and optimized by pilot tests. The computation programs of
membrane manufacturers already include appropriate empirical values from existing
plants. The algorithms contained in the relevant software make it possible to
determine and optimize the design/operating data of the RO process for different
feed and operating conditions (see Sect. 5.4). However, if there are strict quality
requirements with regard to certain substances in the seawater, such as boron and
bromide, it can be necessary to conduct tests for their membrane rejection under real-
world conditions in the desalination stage of a pilot plant.
As already explained in detail in connection with the sampling and analysis
program, it is important that the operation of a pilot plant should, if possible,
cover the full spectrum of variations in seawater quality with regard to their influence
on operating performance. As far as the length of pilot plant testing is concerned, this
means that the plant should be in continuous operation, especially at times of
substantial variation and reduction in water quality. The total length of testing and
the operating times of the pilot plant should be planned accordingly. Consequently,
pilot plant testing can very well extend to a period of 12 months or longer.
The pretreatment and desalination stages of a pilot plant in their design and
configuration should be comparable to those of the projected large-scale plant,
and, like the large-scale plant, they should be capable of continuous operation.
Also, the design parameters should be equivalent to those of the intended SWRO
plant, although, for purposes of optimization, they should be capable of variation.
The membrane desalination stage of the pilot plant should employ the same kind
of membrane elements as those intended for use in the later SWRO plant. The RO
process should have the same membrane configuration as that planned for the large-
scale plant. As far as the configuration of the pressure vessels is concerned, however,
the number of elements can be reduced (three to four elements per pressure vessel
instead of six to eight). Despite this, the flow rates of the individual elements should
more or less correspond to the design data of the large-scale plant. The capacity of
4.2 Basic Design Parameters and Conditions 225

the pilot plant must be capable of satisfying the above-mentioned requirements with
regard to design and operation.
The type and location of seawater extraction for the pilot plant should most
closely correspond to the conditions that will later prevail at the projected SWRO
plant.
So that the product quality achieved by the pilot test can be correlated with the
seawater feed conditions at the pilot plant, and also to make it possible to draw
conclusions about how the proposed large-scale plant will later respond to variations
in seawater quality, the previously initiated sampling and analysis program should
be continued in parallel with operation of the pilot plant. However, the program can
be modified with regard to the number and type of samples and analyses according to
the online measurements in the pilot plant.
In addition to manual sampling and analysis, the following online measurements
are advisable for the pilot plant:
In the seawater feed:

• Temperature
• Conductivity
• pH value
• Turbidity
• SDI or MFI measurement (online or manual)
• Fluorometer for determining the concentrations of chlorophyll and biomass

After the pretreatment stage:

• Turbidity
• Temperature
• SDI or MFI measurement (online or manual)
• pH value
• Particulate count (if UF or MF used for pretreatment)

In the product water after the desalination stage:

• pH value
• Conductivity

These online measurements for seawater extraction and their documentation


make it possible, in addition to the information obtained from the sampling and
analysis program, to produce extensively continuous seawater profiles especially for
temperature, salinity, pH value, as well as particulate- and biomass-related quality at
the planned extraction site. The online measurements after the pretreatment and
desalination stages make it possible to establish a relationship between treatment
efficiency and seawater quality not only with respect to time but also with regard to
important weather-related or seasonal events.
226 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

This also allows the required operating parameters for achieving the desired
treatment efficiency of the pilot plant under the prevailing feed conditions to be
correlated with the respective seawater quality profiles. Consequently, the pilot tests
can provide important information not just for the design of the SWRO plant but also
for its subsequent operation.
More particularly, the results from operation of the desalination stage of the pilot
plant in relation to the residual seawater fouling potential after pretreatment provide
important findings for subsequent operation of the proposed large-scale plant. The
scale and type of fouling (which can be determined, for example, by an autopsy on
the membrane elements), in combination with membrane cleanings in the pilot plant
and their achieved effect, can serve to draw conclusions on the frequency of
chemical membrane cleanings and the required type of chemicals during subsequent
operation of the proposed large-scale plant. So the pilot test results also allow more
accurate calculation of the proportion of operating costs that will be due to mem-
brane fouling.
The subsequent large-scale plant should include similar possibilities for mem-
brane testing to those employed at the pilot plant. For example, after appropriate
modification, the test systems from the pilot plant could be transferred to the large-
scale plant. This would make it possible to test specific membrane cleanings for their
effectiveness and influence on the RO membranes before they are implemented in
the large-scale plant. Also, such a system could serve for testing membranes of the
plant at standard conditions with NaCl solution, verifying their actual performance in
comparison with the permeate capacity and salt rejection of the membrane elements
as specified by the membrane manufacturer.
Especially if the goal is to simulate to the maximum possible extent the seawater
extraction and operating conditions of the proposed large-scale plant, pilot plant
testing will involve a not inconsiderable expenditure of both time and money.
Despite this, in the case of a large-scale SWRO plant, the costs will amount to just
a few percent of total capital expenditure. On the other hand, the optimization results
from pilot testing can lead to appreciable cost savings in connection with the
construction and operation of, for example, the pretreatment stage. In the case of
smaller plant capacities, the relationship between the costs of pilot testing and the
construction costs of the large-scale plant will be considerably less favourable, in
which case it should be considered whether to limit the investigations and tests to
sampling and analysis programs in combination with bench-scale tests and to design
the pretreatment stage with a higher safety margin, i.e. to provide a more extensive
pretreatment process than is warranted by the available information on seawater
quality.
4.2 Basic Design Parameters and Conditions 227

4.2.1.2 SWRO Plant and Plant Systems Capacity and Capacity Range

4.2.1.2.1 Net Output Capacity and Gross Production Capacity


The basis for determining the capacity of an SWRO plant is the net output capacity
CNO, i.e. the product output to be made available by the plant at the interface to the
consumer’s water offtake and distribution system.
However, a desalination plant also has its own internal water consumption, which
must be produced, in addition to the exported capacity, by the plant itself and its
treatment stages, i.e. reverse osmosis desalination, pretreatment, and post-treatment.
Where the water for internal consumption is taken from within the plant will depend
on the quality requirements of the intended use. Such types of water consumption in
an SWRO plant are:

• Low-salinity water for chemical preparation, preparation water/dilution water,


carry water for dosing of chemicals in the pretreatment, reverse osmosis, post-
treatment, and wastewater treatment stages
• Low-salinity water for chemical preparation for chemical cleaning, membrane
preservation, and membrane flushing water
• Pretreated seawater for flushing and concentrate displacement from the mem-
brane systems prior to chemical cleaning and preservation

Normally, the desalinated water for chemical dosing and chemical membrane
cleaning is taken from downstream of the reverse osmosis desalination stages.
Sometimes, however, post-treated product water is used. Pretreated seawater,
which is required for internal water consumption in the pretreatment filtration stages,
is taken from downstream of the pretreatment stage.
This internal water consumption must be taken into consideration in the design of
the SWRO plant and its treatment systems. Therefore, a distinction is made between
the above-mentioned net output capacity CNO and the gross production capacity
CGP, which is calculated according to Eq. 4.3 from CNO and the internal product
water consumption CP,internal for the entire plant or for the treatment system in which
the internal water consumption is produced.
The capacity of an SWRO plant is usually given in m3/d. In the following,
therefore, this is the dimension to which “capacity” always refers. A further distinc-
tion must be made between capacity C in m3/d, product flow F in m3/h, and product
amount Q in m3.

C GP ¼ C NO þ CP,internal ð4:3Þ

CGP ¼ gross production capacity [m3/d]


CNO ¼ net output capacity [m3/d]
CP,internal ¼ internal product water consumption of plant or system [m3/d]
228 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

The water consumption of the chemical dosing systems is largely uniform and
proportional to the capacity at which the desalination plant is operated. For chemical
membrane cleaning, on the other hand, the amount of water required for cleaning is
drawn off intermittently, and the amount consumed for this purpose depends on the
frequency and scale of membrane cleanings. The internal consumption CP,internal of
desalinated water for the above-mentioned operation and maintenance purposes in
an SWRO plant is approximately 1–3% of the net output capacity CNO.

4.2.1.2.2 Availability and Design Capacity


A planned or unplanned outage of a desalination plant results in a loss of capacity,
with the consequence that the net output capacity CNO is sometimes unavailable or
only partially available. During certain periods of operation, there is a reduced output
capacity or effective net output capacity, which is characterized by the plant avail-
ability Ap during this period according to Eq. 4.4.

Ap
CNO,eff ¼ CNO  ð4:4Þ
100

CNO,eff ¼ effective net output capacity [m3/d]


AP ¼ plant availability [%]

Planned outages are required for chemical membrane cleanings, partial mem-
brane replacement, or refurbishment of the membrane processes with new membrane
elements or modules. In addition, there are planned inspections, preventive mainte-
nance measures, and scheduled repairs of the mechanical, electrical, and I&C-related
systems of the SWRO plant.
Unplanned outages are the result of spontaneously occurring defects on the
components, component groups, or systems of the desalination plant and can lead to
unforeseeable outages or capacity reductions of the treatment processes.
With seawater desalination plants or surface water treatment in general, losses of
capacity are also possible as a consequence of deterioration in the quality of the feed
water to the treatment plant with impacts on operational reliability or also product
quality. Such adverse effects on operation or deteriorations in water quality can
cause the capacity of the plant to be reduced or, in the worst case, can make it
necessary for the entire plant to be shut down. Such events must also be categorized
as unplanned outages or capacity reductions. However, such consequences can also
result from quality defects in the plant equipment or from poor engineering.

Plant Availability Based on Capacity Loss


Compared with the availability of its components AU and groups or systems AG,S, the
benchmark of the reduction in plant availability Ap of a desalination process is not
the number of outages in a certain time period, but the reduction in output ΔQoutðτo Þ
experienced by the plant as a result of outages or operation at reduced capacity. To
calculate the plant availability ApðτO Þ based on capacity, the water amount QNO,eff ðτo Þ
4.2 Basic Design Parameters and Conditions 229

effectively produced in a certain operation period t0 is to be compared with the


nominal water amount QNOðτo Þ that the plant should have produced in that operation
period. This makes it possible to determine the output loss ΔQoutðτo Þ : The plant
availability ApðτO Þ can be calculated from the ratio of the effective water amount
QNO,eff ðτo Þ or water losses ΔQoutðτo Þ to the nominally produced water amount QNOðτo Þ.
Therefore, the plant availability based on capacity loss is calculated according to
Eqs. 4.5, 4.6, and 4.7 as

QNO,eff ðτo Þ  100


Apðτo Þ ¼ ð4:5Þ
QNOðτo Þ

QNO,eff ðτo Þ ¼ QNOðτo Þ  ΔQoutðτo Þ ð4:6Þ


 
QNOðτo Þ  ΔQoutðτo Þ  100  ΔQoutðτo Þ

ApðτO Þ ¼ ¼ 1  100 ð4:7Þ
QNOðτo Þ QNOðτo Þ

Apðτo Þ ¼ plant availability during operation period τo [%]


o¼ operation period of plant [hour]
QNO,effðo Þ ¼ net output amount effective during τo [m3]
QNOðo Þ ¼ net output amount nominal during τo [m3]
ΔQoutðo Þ ¼ output loss during τo [m3]

The total output loss ΔQoutðτo Þ during operation period τo is the sum of the
individual output losses ΔQout,iðτout Þ from the number of outages during that operation
period (see Eq. 4.8).

X
i¼z
ΔQoutðτo Þ ¼ ΔQout,iðτout Þ ð4:8Þ
i

ΔQout,iðτout Þ ¼ output loss during outage event i in τo [m3]


i ! z ¼ number of outage events i to z during τo [No/τO]

These outages can be of different duration τout and, during this time, the plant can
have different reductions in capacity or may have been shut down entirely.
Depending on the causes of the outages, the individual output losses ΔQout,iðτout Þ
can be assigned to the planned and unplanned outages of the SWRO plant.
According to Eq. 4.9, therefore, the total output loss ΔQoutðτo Þ is made up of the
output losses from planned and unplanned outages.
230 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

X
i¼z
ΔQoutðτo Þ ¼ ΔQout,iðτout Þ
i

X
i¼z X
i¼z
¼ ΔQout,iðτout Þ,planned þ ΔQout,iðτout Þ,unplanned ð4:9Þ
i i

ΔQout,iðτout Þ,planned ¼ output loss during planned outage event i in τo [m3]


ΔQout,iðτout Þ,unplanned ¼ output loss during unplanned outage event i in τo [m3]

Unplanned outages can be due to component defects. Alternatively, they can be


caused by an unforeseeable deterioration in the quality of the seawater feed to the
plant or by insufficient product quality. In either case, it may be necessary to reduce
the plant capacity or even to shut down the plant. Accordingly, the individual output
losses from unplanned outages ΔQout,iðτout Þ,unplanned must be assigned to these two
causes (see Eq. 4.10).

X
i¼z X
i¼z
ΔQout,iðτout Þ,unplanned ¼ ΔQout,iðτout Þ,unplanned,U
i i

X
i¼z
þ ΔQout,iðτout Þ,unplanned,Q ð4:10Þ
i

ΔQout,iðτout Þ,unplanned,U ¼ component caused output loss during unplanned outage


event i in τo[m3]
ΔQout,iðτout Þ,unplanned,Q ¼ quality caused output loss during unplanned outage event i in
τo [m3]

The output loss ΔQout,iðτout Þ from each individual outage is calculated from the
nominal water amount QNOðτout Þ that the plant would have produced during the
outage period and the amount QNO,eff ðτout Þ actually produced. If the capacity of the
SWRO plant is given in m3/d, the output loss is calculated from the difference of the
nominal and effective flows FNO and Feff and the length of the outage time τout,i.
according to Eq. 4.11.

τout,i
Qout,iðτout Þ ¼ QNOðτout Þ  QNO,eff ðτout Þ ¼  ðC NO  C eff,i Þ
24
¼ τout,i  ðF NO  F eff Þ ð4:11Þ

QNOðτout Þ ¼ net output amount nominal during τout [m3]


QNO:eff ðτout Þ ¼ net output amount effective during τout [m3]
4.2 Basic Design Parameters and Conditions 231

FNO ¼ flow nominal during τout [m3/h]


Feff ¼ flow effective during τout [m3/h]
τout, i ¼ duration time of outage event i [hour]

Calculation of the output losses from planned outages requires knowledge of the
number of outages i ! z in operation period tO and their respective durations tout,i.
Both parameters can be determined during the planning process of an SWRO
plant from:

• Empirical values for the frequency of chemical membrane cleanings


• The design life cycle of the membranes and, therefore, the frequency of mem-
brane replacement and the duration thereof
• The scope and duration of necessary inspections and maintenance measures on
the process equipment

Failure Rate and Operational Availability of Groups and Systems


In terms of plant design, the process technology of an SWRO plant can be assigned
to different levels (see Fig. 4.3).
The frequency and duration of unplanned outages depend primarily on the failure
rates of individual components and on the effect that component outages have on the
groups and systems of the plant and, ultimately, on the plant as a whole in terms of its
availability.
To determine the output losses from unplanned outages, the number of outages zU
caused by an individual plant component is calculated from its specific failure rate
λU. The specific failure rate, in turn, is the reciprocal value of the mean time between
failures MTBFU of a component or unit (Eq. 4.12).

Fig. 4.3 Process technology


levels

Plant

Systems

Groups

Components/ Units
232 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

1
λU ¼ ð4:12Þ
MTBF U

λU ¼ failure rate of unit [F/h]

The duration of an outage τout, i caused by a component defect is known also as


MDTU (mean downtime of unit). This time period includes the necessary repair time
for the component as well as any other outage times (e.g. of an organizational or
administrative nature).
The total outage time of a component or unit τout,U ðτo Þ or MDTU during operation
period τO is then given by the sum of all MDTU and is calculated according to
Eq. 4.13.
X X
MDTU ¼ τO,U  MTBFU ¼ τout,U ðτo Þ ð4:13Þ

MTBFU ¼ mean time between failures for unit [h]


τO,U ¼ operation period of unit [h]
MDTU ¼ τout,U ðτo Þ ¼ mean downtime for unit [h]

Calculation of the output loss from unplanned outages requires knowledge of not
only τout,Uðτo Þ but also the number of outages zU. This parameter is calculated from
failure rate λU and the length of operation period τO,U of the component according to
Eq. 4.14.

zU ¼ λU  τO,U ð4:14Þ

zU ¼ number of outage events of unit during to [No/tO]

Failure rate λU and parameters MTBFU and MDTU are calculated during actual
plant operation by a statistical analysis of outages and by assigning their causes,
frequency, and duration to specific components, groups, or systems. The above-
mentioned parameters, therefore, are of a statistical nature and are, moreover,
specific to certain processes and systems. They are dependent on the quality and
reliability of the component and on the load on the systems during the operation
period as well as on the scope and quality of the maintenance measures designed to
ensure the operability of the component. Other important elements of parameter
MDTU are the scope and quality of spare-parts stocking and the level of efficiency in
terms of the organization and execution of repairs.
Consequently, the values for component outage time and outage frequency
calculated from the basic parameters λU, MTBFU, and MDTU are, like the basic
data from which they were calculated, of a probabilistic nature.
4.2 Basic Design Parameters and Conditions 233

There are tables and databases that contain failure rates as well as repair and
outage times for mechanical, electrical, and electronic components [11–14]. For a
specific process such as an SWRO plant, however, such data can serve only as a
guide, because they will not normally have been determined under the same
conditions, such as type of medium, temperature, pressure, material, and output, as
in an SWRO process. Therefore, reliable data for calculating the failure probability
of components and their impact on the availability of membrane desalination
processes should preferably come from statistical records of operating data over
long operation periods of such plants. Nevertheless, the thus determined parameters
will still exhibit plant-specific particularities, because SWRO plants can differ
considerably in terms of the quality of their equipment and maintenance and also
with regard to their operating conditions.
With knowledge of the failure rates zU and outage times of the components during
operation period τO,U, it is possible to calculate the thereby caused output loss
according to Eqs. 4.8, 4.9, 4.10, and 4.11. However, not every component outage
actually results in a reduction in plant output. If the component belongs to a
redundant group or if the group is part of a redundant system, then there will be
no loss of capacity or only a very small one. How defects and consequent unplanned
outages at component level impact on the higher group and system levels of a plant
must be determined by appropriate availability studies, e.g. using fault tree analysis
or reliability block diagrams [15, 16].
Using the values for MTBFU and MDTU as well as failure rate λU or number of
outages of a component zU, it is possible to estimate the operational availability AU,O
in a given operation period τO,U (Eqs. 4.15, 4.16, 4.17, and 4.17a).
P
MTBFU
f AU,O ¼ P P ð4:15Þ
MTBFU þ MDTU
X
f AU,O ¼ 1  λU  MDTU ¼ 1  λU  τout,U ðτo Þ ð4:16Þ
τout,U ðτo Þ
f AU,O ¼ 1  zU  ð4:17Þ
τO,U

AU,O ¼ f AU,O  100 ð4:17aÞ

fAU,O ¼ factor of operational availability of unit []


AU,O ¼ operational availability of component or unit [%]

The operational availability AU,O is defined as the probability with which a


component will operate in a given operation period under the existing operation
and maintenance conditions.
Table 4.3 Failure rates and availability of mechanical equipment components
234

Mean time Mean Outage events of Availability of Unavailability factor


between failures downtime Failure rate component per year component of component
4

MTBFU MDTU λU zU AU fUAU


No. Component/unit [h] [h] x 106 [F/h] [No/year] [%] x 104 []
1 Centrifugal pumps
• Seawater feed, 40,000 24 25.0 0.22 99.940 5.996
low/medium pressure
• RO 1st pass, high 30,000 24 33.3 0.29 99.920 7.994
pressure
• RO 2nd pass, medium 50,000 16 20.0 0.18 99.968 3.199
pressure
2 Membranes and pressure 100,000 8 10.0 0.09 99.992 0.800
vessels
3 Energy recovery systems 35,000 16 28.6 0.25 99.954 4.569
4 Motors of centrifugal 100,000 8 10.0 0.09 99.992 0.800
pumps
5 Dosing pumps
• Neutral and alkaline 60,000 4 16.7 0.15 99.993 0.667
media
• Acids 30,000 8 33.3 0.29 99.973 2.666
6 Automatic valves 30,000 8 33.3 0.29 99.973 2.666
7 Control valves 50,000 8 20.0 0.18 99.984 1.600
8 Level control 40,000 4 25.0 0.22 99.990 1.000
9 Online analyser
• Meters (conductivity, 50,000 4 20.0 0.18 99.992 0.800
turbidity, etc.)
Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .
• Analysers (pH, ORP, 20,000 8 50.0 0.44 99.960 3.998
SDI, oil, etc.)
10 Flow meter 150,000 4 6.7 0.06 99.997 0.267
11 Piping, high pressure 100,000 24 10.0 0.09 99.976 2.399
12 Agitators 100,000 16 10.0 0.09 99.984 1.600
13 Air compressors 50,000 8 20.0 0.18 99.984 1.600
14 Transformers 400,000 30 2.5 0.02 99.993 0.750
4.2 Basic Design Parameters and Conditions
235
236 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 4.3 contains guide values for MTBF, outage time MDT, and failure rate λ
for various process components of an SWRO plant and the therefrom calculated
values for the number of annual outages zU, operational availability AU,O, and
operational unavailability fUAU,O (Eq. 4.18).
The data should be understood as average values for seawater desalination plants
with state-of-the-art equipment and materials and a high standard of maintenance.
The MDT values for outage times apply to defects that can be remedied by spare
parts, i.e. defects that do not require the replacement of components (no component
breakdown).
f UAU,O ¼ 1  f AU,O ð4:18Þ

fUAU,O ¼ factor of operational unavailability of unit []

The components, groups, and systems at the process levels of a plant are usually
serially connected within those levels and also between lower levels and higher
levels. More especially, there are serial connections between the treatment systems
in a process chain: in an SWRO plant, for example, between the seawater extraction
systems, the pretreatment and reverse osmosis desalination processes, and the post-
treatment stage. As far as operational availability in the case of serial connection is
concerned, the overall availability AG,S is calculated by multiplying the individual
availabilities Ai,U,O of the individual groups or systems (see Eqs. 4.19, 4.19a, and
4.19b).

iY
=n
f AG,S,O = f AUi,O ¼ f AU1,O  f AU2,O  . . . f AUn,O ð4:19Þ
i=1

AG,S,O ¼ f AG,S,O  100 ð4:19aÞ

AUi,O ¼ f AUi,O  100 ð4:19bÞ

fAG,S,O ¼ operational availability factor of group or system []


AG,S,O ¼ operational availability of group or system [%]
AUi,O ¼ operational availability of component or unit [%]

In such a serial chain, therefore, the overall availability is crucially determined by


the link with the lowest availability. Thus, to achieve a high overall availability, the
individual availabilities along the process chain must be at least equal and, if
possible, equally high.
For a single train of the reverse osmosis desalination system of an SWRO plant
consisting of serially connected booster pump, high-pressure pump, energy recovery
system, and RO membrane unit with the associated pipes and valves, it is possible
according to Eq. 4.19 and with the data from Table 4.3 to estimate the system
availability AAG,S at approximately 99.6–99.8%. However, this availability is a
statistical, mean annual value. If, at a given time, there is an unplanned outage of
4.2 Basic Design Parameters and Conditions 237

one of the reverse osmosis trains, e.g. because of several component defects, this can
result for a short time period in a considerable loss of output capacity of the overall
plant on account of the significant downtime. Therefore, systems with capacity-
critical components with a high downtime in the event of an unplanned outage must
therefore be configured and maintained during the planning of an SWRO plant in
such a way that plant availability is not affected adversely by such events.
Possible ways of safeguarding and increasing the availability of the plant include:

• When selecting the process components, to pay attention to available experience


with the units in identical or similar applications as well as to high quality (with
regard to the choice of materials in order to prevent corrosion in the case of a
seawater desalination plant) and ease of maintenance
• To design components, process groups, and process systems with appropriate
levels of redundancy, so that, in case of planned or unplanned outages, the plant
can continue to operate without loss of capacity, thanks to the existence of
corresponding standby systems
• In the case of planned outages, to reduce the loss of output capacity by reducing
the duration and frequency of outages, e.g. by grouping different maintenance
activities within the same time period, as well as by increased use of materials and
increased expenditure on human resources
• To so design the capacity of the plant that capacity losses can be offset before or
after their occurrence by operating the plant at increased capacity

Redundancy of Components and Systems


Redundancy is used in process plants, especially at the component level, where
several components in a group are redundantly interconnected in parallel with
appropriate backup units, e.g. several pumps forming a single pump group. Redun-
dancy is also built into the system level of a process plant, albeit usually on a smaller
scale, because redundancy at the system level is considerably more expensive than at
the component/group level. Normally, there is a redundancy of n + 1, where the
nominal capacity of a group or system is divided between n components and one
element with identical capacity is provided as a backup unit (see Eqs. 4.23 and 4.24).
In the case of the reverse osmosis process of an SWRO plant, an example of
redundancy at the system level is to design the process with standby trains or to
split the pretreatment systems into several lines. Redundancy focuses on those
components, groups, or systems whose failure would result in an outage of the
desalination plant and, therefore, in a total loss of its capacity.
According to the redundant configuration of the group or system, i.e. the total
number of components and the number of operational and standby components, in
the event of the failure of one or more of them, the group/system will still be capable
of operation at partial or full capacity. The still available capacity Feff of a redundant
component group or system in the event of an outage and the capacity loss during the
outage time can be calculated according to Eqs. 4.20 and 4.21 using a redundancy
factor fRed,i.
238 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

F eff ¼ F NO  f Red,i ð4:20Þ


 
ΔQout,iðτout Þ ¼ τout,i  F NO  1  f Red,i ð4:21Þ

FNO ¼ flow nominal during tout [m3/h]


Feff ¼ flow effective during tout [m3/h]
τout, i ¼ duration time of outage event i [hour]
fRed,i ¼ redundancy factor of outage event i []

The redundancy factor results from the configuration of the components in the
group or system and their assignment to operation and standby mode as well as from
the actual number of components that are out of operation (see Eqs. 4.22, 4.23, and
4.24).

CD n  nsb  nout
f Red,G,S ¼ ¼ total ð4:22Þ
C NO nop

CD ¼ C U  ðntotal  nsb  nout Þ ð4:23Þ


C NO
nop ¼ ð4:24Þ
CU

CD ¼ deliverable capacity of group or system [m3/d]


CNO ¼ nominal capacity of group or system [m3/d]
CU ¼ selected capacity of units in group or system [m3/d]
ntotal ¼ total number of units of group or system []
nop ¼ number of units to be operated for nominal capacity in group or system []
nsb ¼ number of units actually on standby in group or system []
nout ¼ number of units actually in outage in group or system []
fRed,G,S ¼ redundancy factor of group or system in operation and standby configura-
tion and outage situation []

By forming redundant groups and systems, it is possible to considerably improve


their availability fAG,S,O in comparison with the availability fAU,O of the individual
component and also to reduce the multiplication effect of availability between one
process level and the next higher level (see Eq. 4.25).
 n þ1
f AG,S,O ¼ 1  f  1  f AU,O sb ð4:25Þ
 
nop þ nsb !
f ¼  
nop  1 !  ðnsb þ 1Þ!

Depending on the degree of redundancy of its systems and the quality of its
components, the average annual availability of an entire SWRO plant can be
4.2 Basic Design Parameters and Conditions 239

between 93% and over 98%, equivalent to a capacity reduction of the plant by
between 2% and 7%. If the plant is suitably provided with redundant groups and
systems, the proportion of capacity reduction due to unplanned outages (excluding
external influences) will be around <1% to max. 2%.

Plant Design with Increased Design Capacity


Another way to prevent capacity losses as a consequence of outages is to design the
plant with a capacity that is higher than the net output capacity CNO. In the event of a
planned outage, the plant can then be operated at the higher capacity for a certain
period of time until the capacity loss caused by the outage has been made good. The
same is possible after an unplanned outage, in which case, however, either the plant
or the consumer must have an appropriate storage capacity for storing the increased
output prior to delivery into the distribution network.
With this kind of design, the capacity used as the basis for calculating the
individual plant systems is called the net design capacity CND, and the underlying
plant availability is called the design availability APD (Eq. 4.26). These can then be
used to calculate the gross design capacity CGD, which includes the internal water
consumption CP, internal of the plant (Eq. 4.27).

C ND¼CNO 100 ð4:26Þ


APD

C GD ¼ C ND þ C P,internal ð4:27Þ

CND ¼ net design capacity [m3/d]


CGD ¼ gross design capacity [m3/d]
APD ¼ design availability [%]
CP,internal ¼ internal water consumption capacity [m3/d]

In case of a planned outage, for example, the duration of a certain maintenance


activity can serve as a basis for calculating the design availability APD of the plant.
Conversely, for unplanned outages, the maximum repair time for a given component
or system can be used. If the expected availability AP of the plant is known from an
availability analysis of the SWRO plant, then this value can be used as a basis to
calculate the net design capacity APD.
With regard to a capacity loss of an SWRO plant due to deterioration in seawater
feed quality or because of quality problems on the product side, it is normally
impossible to predict either the timing of the event, its frequency, or the duration
of the capacity loss or possible plant outage. Although the chosen pretreatment
process of an SWRO plant is designed to maintain the seawater feed quality to the
desalination stage across the entire seawater quality range, heavy biological growth
in the sea (algal bloom, jellyfish bloom, etc.) or another suddenly occurring deterio-
ration in quality (e.g. oil spill) can be too much for even a correctly designed
pretreatment stage to handle. In such a case, if the desalination plant has been
designed with an appropriate capacity reserve, i.e. an increased net design capacity
240 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

CND, this makes it possible to offset production losses caused by such events by
subsequently operating the plant at an increased capacity.
The design capacity of an SWRO plant will depend also on the intended mode of
operation as well as on the capacity availability required by the consumer. If the
desalination plant is part of a water supply network in which capacity losses can be
predictably offset from other sources, then it will not be necessary to provide high
degrees of redundancy, e.g. at system level (such as backup trains in the reverse
osmosis system). In this case, the plant can be temporarily operated according to
Eq. 4.4 with an effective net output capacity CNO,eff caused by capacity losses during
chemical membrane cleaning, membrane replacement, or maintenance activities. If,
however, the plant is crucial to the basic supply to a consumer and its net output
capacity CNO needs to have an almost 100% availability, then the plant should be
designed not only with a high redundancy at both the component and system levels
but also with a capacity reserve based on a net design capacity CND according to
Eq. 4.26 and a conservative plant availability APD.

4.2.1.2.3 Capacity Range of Plant: Minimum and Maximum Output Capacity


and Overload Capacity
The extent to which the output capacity of an SWRO plant is variable can be
determined both by the flexibility of the product capacity of the actual membrane
units and by the modular design of the reverse osmosis membrane process.
If a high flexibility of the SWRO plant is required, this must be appropriately
taken into consideration in the design of the membrane elements of the reverse
osmosis trains and/or by their modular configuration.
The range within which the output capacity of a membrane train can be varied is
primarily determined by the design and operating criteria of the membrane elements
and modules used in the reverse osmosis desalination systems of the plant. These
criteria are specified by the membrane manufacturer and are included in the calcula-
tion algorithms in the design programs of the membrane supplier (see Sects. 5.2 and
5.2.1). These parameters control the flow conditions in the membrane elements in
such a way that excess concentrations at the surface of the separation membranes are
avoided as far as possible. For this purpose, it is necessary that a sufficiently high
flow velocity is available at the membrane surface, but at the same time the permeate
flow through the membrane must be in the right ratio to its overflow.
As far as membrane design is concerned, the design framework for the reverse
osmosis system is mainly determined by the choice of a certain product recovery rate
(see Sect. 4.2.1.3) and a specific permeate flux (see Sects. 5.2 and 5.2.1). These
parameters for the design of the plant can be based on the maximum values within
the range of guide values specified by the membrane manufacturer, or they can also
be more conservatively selected.
If it is desired that the product recovery rate should be as high as possible (i.e. a
product recovery rate of around 45% or higher) and if this product recovery rate is to
be maintained during operation of the plant, then the product capacity of a reverse
osmosis train with spiral wound elements can be reduced by max. approx. 10% while
remaining within the allowable limits for the membrane elements. If the plant is
4.2 Basic Design Parameters and Conditions 241

designed with a more conservative product recovery rate (around 40%), it is possible
for an RO train to be operated with a product capacity approx. 20% below the design
capacity. Then with the lower capacity rate, the operating pressure of the plant is
reduced accordingly.
If an RO train is configured in such a way that, to achieve a given capacity
reduction, its product recovery rate can be significantly lowered during operation,
then, with regard to membrane design, it is possible to achieve minimum capacities
of up to 30–40% of the design capacity. Depending on the flow characteristics of the
membrane elements, this requires a reduction of the product recovery rate in the
range of up to 10–15% of the design value. A reduction of the product recovery rate
also is accompanied by a reduction of the required operating pressure for the
membranes, and the feed pumps must be appropriately designed for this mode of
operation in terms of their characteristics and type of control.
An increased capacity or overload of the membrane process is accomplished by
raising the operating pressure in the individual trains. This causes an increase in
permeate flux of the membranes and consequently an increase in product output. In
this case, the maximum possible output is limited by the manufacturer’s
specifications for the maximum allowable specific permeate flux and the maximum
allowable operating pressure of the membranes (see Sects. 5.2 and 5.2.1). The degree
of overload at which the maximum allowable specific permeate flux is reached will
depend on how conservatively or progressively this design parameter was chosen.
Normally, with a conservative design, the design capacity of a train can be exceeded
by max. 10–20%.
If the desired flexibility of the membrane process is to be achieved by suitable
configuration of its modular construction, the total capacity of the system is split
between the capacities of the individual trains of the reverse osmosis process, so that,
to attain an appropriate capacity reduction, one or more modules/trains can be shut
down or, conversely, to run an appropriate overload, one or more modules/trains can
be started up. In this case, for example, the capacity of an individual train is designed
to correspond to the desired lowest capacity of the desalination process. When the
RO process is operated at this capacity, only one train will be in operation to deliver
the minimum load. Therefore, assuming a product capacity of an RO seawater
desalination system of 100,000 m3/day and a desired minimum output of 20% of
that capacity, this means a train output of 20,000 m3/day. In this case, it is possible
for the individual trains to be operated at a constant product recovery rate according
to the nominal design.
This mode of operation can result in additional operating costs for chemicals as
well as higher in-process consumption of water and energy due to the fact that if
individual trains are shut down for a longer period, they need to be conserved with
chemicals after having previously undergone chemical cleaning. This leads, of
course, to the consumption of not only chemicals but also in process water and
energy (see Sect. 4.2.3.3; Chap. 5; Sects. 5.5 and 5.5.2.5).
All flexibility, overload, and low load mode of operation requirements for the
SWRO plant also need the configuration and modular design of seawater extraction,
242 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

pretreatment, post-desalination, and product post-treatment to be appropriately con-


sistent with the RO seawater desalination system design and operating conditions.

4.2.1.3 Product Recovery Rate and RO Membrane System Basic


Equations
The product recovery rate or product recovery coefficient YRO is a basic design
parameter of a reverse osmosis system. Once YRO has been specified, this determines
the required feed capacity CF,RO of the membrane system for producing the desired
product/permeate capacity CP,RO. YRO is calculated as the quotient of product
capacity CP,RO and feed capacity CF,RO or product capacity and concentrate capacity
CC,RO of the plant according to Eqs. 4.28, 4.29, and 4.30. The feed capacity CF,RO to
the desalination stage is then determined according to Eqs. 4.31 and 4.32.

CP,RO
Y RO ¼ ð4:28Þ
C F,RO

C F,RO ¼ CP,RO þ C C,RO ð4:29Þ


CP,RO
Y RO ¼ ð4:30Þ
C P,RO þ C C,RO

C P,RO ¼ CF,RO  Y RO ð4:31Þ


C P,RO
C F,RO ¼ ð4:32Þ
Y RO

YRO ¼ recovery coefficient RO []


CP,RO ¼ product capacity RO [m3/d]
CF,RO ¼ feed capacity RO [m3/d]
CC,RO ¼ concentrate capacity RO [m3/d]

Based on the mass and concentration balance presented in Fig. 4.4, it is possible
to derive further basic equations for calculating the part streams of a reverse osmosis
unit and their concentrations.
Thus, the concentrate capacity of the RO unit is calculated from its feed capacity
CF,RO or product capacity CP,RO and YRO according to Eqs. 4.33 and 4.34.

C C,RO ¼ CF,RO  ð1  Y RO Þ ð4:33Þ


 
1
C C,RO ¼ C P,RO ∙ 1 ð4:34Þ
Y RO

Drawing up the concentration balances for the part streams of an RO unit requires
knowledge of not only the feed concentration cF,RO to the membrane but also the salt
passage factor SPRO or salt rejection factor RRO. Salt passage factor SPRO and salt
rejection factor RRO are calculated from concentration in product cP,RO and
4.2 Basic Design Parameters and Conditions 243

Reverse Osmosis
RO

CF,RO CP,RO
Feed cF,RO cP,RO Product

CC,RO
cC,RO

Concentrate

Fig. 4.4 Basic balance of an RO membrane unit

concentration in feed cF,RO as shown in Eqs. 4.35 and 4.36. The concentration in
product cP,RO is calculated from the concentration in feed cF,RO and the salt passage
factor SPRO according to Eq. 4.37.

cP,RO
SPRO ¼ ð4:35Þ
cF,RO
cP,RO
RRO ¼ 1  SPRO ¼ 1  ð4:36Þ
cF,RO

cP,RO ¼ cF,RO  SPRO ¼ cF,RO  ð1  RRO Þ ð4:37Þ

SPRO ¼ salt passage factor RO []


RRO ¼ salt rejection factor RO []
cP,RO ¼ concentration in product [mg/l, g/m3]
cF,RO ¼ concentration in feed [mg/l, g/m3]

Once the product recovery rate YRO has been specified, this also determines the
degree of concentration of the salt content and components in the feed to the RO unit
during the desalination process and the resultant composition of the reverse osmosis
concentrate (see Eq. 4.38).

C P,RO cC,RO  cF,RO


Y RO ¼ ¼ ð4:38Þ
C F,RO cC,RO  cP,RO

cC,RO = concentration in concentrate [mg/l, g/m3]

The degree of concentration of a component in the feed or the salt content in the
feed to the desalination process in the membrane unit is defined also by the ratio of
its concentration in the concentrate and its concentration in the feed as concentration
244 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 4.4 RO unit—basic part stream concentration equations


Parameters
No. To be calculated Available Equations
1 cC,RO cF,RO, cP,RO,YRO cC,RO ¼ cF,RO Y RO cP,RO
1Y RO
2 cF,RO,YRO, SPRO RO SPRO
cC,RO ¼ cF,RO  1Y1Y RO

3 cF,RO,YRO, RRO ð1RRO Þ


cC,RO ¼ cF,RO  1Y RO1Y RO

4 cF,RO, CF cC, RO ¼ CF  cF, RO


5 cP,RO,YRO, SPRO 1
Y RO
cC,RO ¼ cP,RO  SP1Y
RO
RO

6 cp,RO, YRO, RRO 1


Y RO
cC,RO ¼ cP,RO  1R1Y
RO
RO

7 cP,RO cF,RO, SPRO cP, RO ¼ cF, RO  SPRO


8 cF,RO, RRO cP, RO ¼ cF, RO  (1  RRO)

factor CF (see Eq. 4.39). If the concentration in product cP,RO can be ignored because
of a very low salt passage factor SPRO or a very high salt rejection factor RRO of the
membrane, then Eq. 4.39 can be simplified to Eq. 4.39a.

cC,RO 1  Y RO  SPRO
¼ CF ¼ ð4:39Þ
cF,RO 1  Y RO
1
CF ffi ð4:39aÞ
1  Y RO

CF ¼ concentration factor []

The product recovery coefficient YRO is calculated from the concentration factor
CF as shown in Eqs. 4.41 and 4.41a.

CF  1
Y RO ¼ ð4:40Þ
CF  SPRO
CF  1 1
Y RO ffi ffi1 ð4:40aÞ
CF CF
Table 4.4 gives the equations for determining the component concentrations in
concentrate cC,RO and in product cP,RO of a reverse osmosis unit from their concen-
tration in feed cF,RO. This assumes knowledge of product recovery coefficient YRO,
salt passage factor SPRO, salt rejection factor RRO, and concentration factor CF.
The product recovery rate is an important parameter for the design and operation
of an SWRO plant. It has a key influence on the dimensioning of the membrane
desalination process and associated systems of the plant as well as on a multiplicity
of operating parameters, such as the in-process consumption of energy and
chemicals. Table 4.5 shows how an increase or reduction of the product recovery
4.2 Basic Design Parameters and Conditions 245

Table 4.5 Impact of product recovery on SWRO design and


operation

Product Recovery
Impact on Design / Operation YRO1
(for constant output capacity) Increase Decrease
of value of value
RO seawater process
Feed
• Capacity
• Pressure
• Chemicals consumption
Concentrate
• Capacity
• Pressure
• TDS
• Energy recovery system capacity
Product - TDS
RO design
• No. of concentrate stages
• Membrane area
Energy consumption
Post desalination (RO 2nd pass)
Feed
• Capacity
• Pressure
• Chemicals consumption
Concentrate
• Capacity
• TDS
Product - TDS
Energy consumption
Seawater extraction and Pre-treatment
Capacity
Chemicals consumption
Energy consumption
Waste water treatment
Capacity
Chemicals consumption
Energy consumption
Concentrate outfall
Capacity
Outfall concentration
246 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

rate YRO1 in the RO desalination process affects the various treatment stages of an
SWRO plant and their operating conditions.
The product recovery rate determines the degree of concentration of the
components in the seawater to be desalinated and, therefore, the mean osmotic
pressure of the resulting concentrate and, consequently, constitutes the basis for
the operating pressure of the reverse osmosis desalination stage. The concentration
of seawater components in the concentrate, in turn, has an influence on the salt
content in the product. The salt content in the product of the membrane desalination
process, in turn, is decisive for the required capacity of a post-desalination stage.
Furthermore, the amount of concentrate produced by the desalination process is
dependent on YRO1, and consequently, this also determines the dimensioning of the
energy recovery systems of the RO and the outfall of the plant. Also, the design of
the seawater extraction and pretreatment systems of the SWRO plant is determined
by the product recovery rate.
The various influences of the product recovery rate on the design and operation of
an SWRO plant (see Table 4.5) mean that the optimization of this parameter during
the planning process for the plant requires a multiplicity of interrelated technical and
economic aspects to be taken into consideration.
Basically, optimization of the product recovery rate covers all the process systems
from seawater extraction through pretreatment, desalination systems with main and,
if required, post-desalination as well as wastewater treatment and outfall of the plant.
The optimization process relates both to the costs of buildings and process equip-
ment and to the operation and maintenance costs of the respective systems and,
therefore, to the production costs of the plant as a whole. It is advantageous for the
optimization process to consider not only the current annual costs at the time the
plant is being planned but also the likely life cycle costs over the lifetime of the plant
(see also [23], Chap. 9 in Volume 2).
Depending on the salt content of the seawater and the requirements with regard to
the design and operating conditions of the SWRO plant, the range of the product
recovery rate is between 35% and 55% for its RO membrane seawater desalination
systems part.

4.2.1.4 Product Water Quality, Post-desalination, and Post-treatment


The requirements of the end consumer(s) with regard to the composition and quality
of the product water from the SWRO plant are important criteria with a significant
influence on the design of the seawater membrane stage as well as on the choice of
necessary post-desalination and post-treatment stages for the reverse osmosis per-
meate. Post-desalination is necessary if the total salt content or the concentrations of
certain components in the permeate of the RO seawater desalination stage are still
above the required target values for the final product water from the SWRO plant.
Apart from the salt content of the product water, also the chloride content can be
subject to limits or, if drinking water is being produced from seawater, a limit can
also be set for other components of seawater, such as bromide and boron.
The requirements with regard to the composition and quality of drinking water are
quite extensive. This is shown by way of example in the parameters listed in No. 5 of
4.2 Basic Design Parameters and Conditions 247

Table 4.1 and is explained in detail in [23], Chap. 3 in Volume 2. In addition to


achieving the target values for TDS, chloride, bromide, and boron by post-
desalination of the permeate from the seawater desalination stage, to achieve the
required concentration values for calcium and alkalinity and also to obtain the guide
values for minimizing the corrosivity of the drinking water, it is additionally
necessary to post-treat the product water from the reverse osmosis systems of the
SWRO plant. This includes the necessary measures for hygienization of the drinking
water before it is delivered to the consumer(s).
In addition to setting the target values for TDS, chloride, bromide, and boron by
appropriate post-desalination of the permeate of the seawater desalination stage, an
additional post-treatment of the product water of the reverse osmosis systems of
SWRO is required to set predetermined concentration values for calcium and
alkalinity and to set the guide values to minimize the corrosiveness of the drinking
water. The necessary measures for the hygienization of the drinking water before it is
delivered to the consumers are also carried out there.

4.2.1.4.1 Capacity Factor and Capacity Design of Post-desalination


As far as drinking water production in an SWRO plant is concerned, it is necessary,
depending on the required composition of the drinking water, to suitably reduce the
total salt content, chloride content, and usually also the concentration values for
bromide and boron in the permeate from the seawater desalination stage in order to
achieve the respective target values in the final product water. This is usually done in
an RO post-desalination stage, which is fed with a certain proportion of the product
water from the first pass of the desalination plant. The other proportion of the
permeate from the first pass bypasses the post-desalination stage, and both part
streams—the product of the post-desalination stage and the bypass water—are
then merged to yield the final product (see Fig. 4.5). To produce drinking water,
the post-desalination stage is normally equipped with brackish water membranes,
i.e. RO membranes that have a lower salt rejection than seawater membranes and are
operated at a lower feed pressure (see Chap. 5, Annex 5.A3, Table 5.39).
The proportion CP,RO2 of the product output from reverse osmosis desalination
CP,M that must be fed to the post-desalination unit RO2 in order to achieve the
required concentration of a constituent in the final product is defined by the capacity
factor fC,RO2. The factor is calculated from the concentration of the relevant constit-
uent in the permeate from the post-desalination stage cP,RO2 and in the product water
from the main desalination stage that bypassed the post-desalination stage (cB,RO2) as
well as its desired concentration in the final product cP,M (see Eq. 4.41).

f C,RO2¼CP,RO2 ¼ cB,RO2 cP,M ð4:41Þ


C P,M cB,RO2 cP,RO2

fC,RO2 ¼ capacity factor RO2 (RO 2nd pass) []


CP,RO2 ¼ capacity of RO2 unit [m3/d]
CP,M ¼ capacity of mixed RO product ≌ CNO,SWRO [m3/d]
248 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

CB,RO2 ¼ concentration of constituent of CB,RO2 [mg/l]


cP,M ¼ concentration of constituent of CP,M [mg/l]
cP,RO2 ¼ concentration of constituent of CP,RO2 [mg/l]

If the post-desalination stage RO2 is both fed with the combined permeate output
from seawater desalination unit RO1 and is also bypassed by the same product, then
the concentration of a constituent in the bypass cB,RO2 will be the same as that in the
feed to RO2 (see Eq. 4.42). The concentration in both part streams is then identical to
the product concentration cP,RO1 of the main desalination stage RO1, and Eq. 4.41
then changes accordingly to Eq. 4.43.

cB,RO2 ¼ cP,RO1 ð4:42Þ


CP,RO2 c  cP,M
f C,RO2 ¼ ¼ P,RO1 ð4:43Þ
C P,M cP,RO1  cP,RO2

cP,RO1 ¼ concentration of constituent of CP,RO1 [mg/l]

If, however, a part stream of lower-salinity permeate is drawn from the membrane
pressure vessels of the main desalination stage RO1 according to the so-called split
partial method (see Sects. 5.2.3 and 5.2.3.2) and is supplied to the bypass, then the
concentration of constituent cB,RO2 will there be lower than in the feed to post-
desalination stage RO2 and fC,RO2 can then be calculated as shown in Eq. 4.41.
With knowledge of the salt passage factors of the membranes of the main
desalination stage RO1 (SPRO1) and post-desalination stage RO2 (SPRO2) for the

Seawater feed CB,RO2


from pre- cB,RO2
treatment
CP,Pr CF,RO1 CP,RO1 CP,RO2 CP,M
RO 1 RO 2
cP,Pr cF,RO1 cP,RO1 CF,RO2 cP,RO2 cP,M

YRO1 YRO2 Final RO


product
CC,RO1 cC,RO1 CC,RO2 cC,RO2

CC,RO1, RO2
RO systems
cC,RO1, RO2 discharge

Pre-treated Seawater
RO product
Concentrate

Fig. 4.5 Basic balance of a two-pass RO system


4.2 Basic Design Parameters and Conditions 249

relevant constituent and its concentration in the feed to the reverse osmosis desali-
nation unit cF,RO1, it is possible to calculate capacity factor fC,RO2 with Eq. 4.44.
With knowledge of the salt rejection factors of the membranes in RO1 and RO2
(RRO1 and RRO2) for the constituent, fC,RO2 is calculated according to Eq. 4.45.

SPRO1  ccF,RO1
P,M

f C,RO2 ¼ ð4:44Þ
SPRO1  SPRO2  SPRO1

SPRO1 ¼ salt passage factor for constituent of membranes in RO1 []


SPRO2 ¼ salt passage factor for constituent of membranes in RO2 []

ð1  RRO1 Þ  ccF,RO1
P,M

f C,RO2 ¼ ð4:45Þ
ð1  RRO1 Þ  ð1  RRO1 Þ  ð1  RRO2 Þ

RRO1 ¼ salt rejection factor for constituent of membranes in RO1 []


SPRO2 ¼ salt rejection factor for constituent of membranes in RO2 []

To determine the concentration factor fC,RO2 of a constituent whose target con-


centration in the final product from the SWRO plant is to be achieved by post-
desalination, this is done with reference to its maximum concentration both in the
product from the main desalination stage and in the permeate from the post-
desalination stage, as is given by the RO membrane calculations (see Sect. 5.4). If
there are target values for several constituents, the capacity factor of the constituent
with the highest value is used as the basis for calculating the necessary capacity of
the post-desalination stage CP,RO2 according to Eq. 4.46.

C P,RO2 ¼ f C,RO2  CP,M ð4:46Þ

If the target values for constituent concentrations are based on the final product
from the SWRO plant, then, if a limit has been specified for the salt content in the
final product, it must be taken into consideration that post-treatment of the product
water to increase its hardness and alkalinity will also increase its salt content by an
amount ΔcTDS,Pt. The required salt content cTDS,CP,M in the mixed product from
desalination systems RO1 and RO2 is then to be lower than the target value in the
final product from the SWRO plant cTDS,NO,SWRO according to Eq. 4.47.

cTDS,CP,M ¼ cTDS,NO,SWRO  ΔcTDS,Pt ð4:47Þ

cTDS,CP,M ¼ concentration of TDS in CP,M [mg/l]


cTDS,NO,SWRO ¼ concentration of TDS in CNO,SWRO [mg/l]
ΔcTDS,Pt ¼ TDS increase during post-treatment [mg/l]
250 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Calculation of the capacity factor for TDS must then be based on the salt content
in the mixed product cTDS,CP,M calculated according to Eq. 4.47 and not on the target
value cTDS,NO,SWRO in the final product from the SWRO plant.
Frequently, the post-desalination stage is designed for a higher capacity than that
given by calculating the capacity coefficients. Such a capacity reserve of RO2 makes
it possible, in order to ensure achieving the required quality of the end product, to
balance out a disproportionate increase in the salt content or in the concentrations of
other constituents in the permeate from the main desalination stage RO1—e.g. as a
result of a deterioration in seawater quality or because of increased membrane
fouling or scaling in RO1 and/or RO2—by operating the post-desalination stage at
an increased capacity.

4.2.1.4.2 Product Recovery of a Two-Pass RO System and SWRO Plant


In a two-pass system of the kind presented in Fig. 4.5, a distinction must be made
between product recovery from the RO main desalination stage (first pass) and that
from the post-desalination stage (second pass). If the membrane process for seawater
desalination is operated at a recovery coefficient YRO1 of, depending on the operating
conditions of the SWRO plant, between 35% and 55%, then the recovery coefficient
YRO2 of the post-desalination stage will normally be between 85% and 95%.
The recovery coefficient of the first pass YRO1 and the respective capacities of its
part streams are calculated as shown in Eqs. 4.48, 4.49, and 4.50.

CP,RO1
Y RO1 ¼ ð4:48Þ
C F,RO1

CP,RO1 ¼ C F,RO1  Y RO1 ð4:49Þ

C C,RO1 ¼ CF,RO1  ð1  Y RO1 Þ ð4:50Þ

CP,RO1 ¼ product capacity RO 1st pass [m3/d]


CF,RO1 ¼ feed capacity RO 1st pass [m3/d]
CC,RO1 ¼ concentrate capacity RO 1st pass [m3/d]
YRO1 ¼ recovery coefficient RO 1st pass []

The recovery coefficient YRO2 and the part stream capacities of the second pass
are calculated in a similar manner (see Eqs. 4.51, 4.52, and 4.53).

CP,RO2
Y RO2 ¼ ð4:51Þ
C F,RO2

CP,RO2 ¼ C F,RO2  Y RO2 ð4:52Þ

C C,RO2 ¼ CF,RO2  ð1  Y RO2 Þ ð4:53Þ


4.2 Basic Design Parameters and Conditions 251

YRO2 ¼ recovery coefficient RO 2nd pass []


CP,RO2 ¼ product capacity RO 2nd pass [m3/d]
CF,RO2 ¼ feed capacity RO 2nd pass [m3/d]
CC,RO2 ¼ concentrate capacity RO 2nd pass [m3/d]

The recovery coefficient YRO1,RO2 of the two-pass reverse osmosis system is


determined from the ratio of its product capacity CP,M and the feed capacity to the
reverse osmosis system CP,Pr according to Eq. 4.54. This equation then also yields
the part stream capacities of the total reverse osmosis process (see Eqs. 4.55 and
4.56).

CP,M
Y RO1,RO2 ¼ ð4:54Þ
C P,Pr

CP,M ¼ C P,Pr  Y RO1,RO2 ð4:55Þ

C C,RO1,RO2 ¼ C P,Pr  ð1  Y RO1,RO2 Þ ð4:56Þ

YRO1,RO2 ¼ recovery coefficient of RO tract (RO1 and RO2) []


CP,M ¼ RO tract product capacity [m3/d]
CP,Pr ¼ treated seawater feed capacity to RO tract [m3/d]
CC,RO1,RO2 ¼ concentrate discharge capacity of RO tract/SWRO plant [m3/d]

The total recovery coefficient YRO1,RO2 of the two interconnected reverse osmosis
stages is lower than that of the first pass of the two-pass system. It is necessary to
distinguish whether the concentrate from the second pass is discharged from there or
is recycled into the feed to the first pass. In the former case, i.e. without concentrate
recycling, the feed capacity CF,RO1 to RO1 is the same as the feed capacity CP,Pr
from the SWRO pretreatment stage to the reverse osmosis system. YRO1,RO2 is then
calculated from the recovery coefficient of the first pass YRO1, the recovery coeffi-
cient of the second pass YRO2, and the capacity factor fC,RO2 for the post-desalination
stage, as shown in Eq. 4.57.

C P,M C Y RO1
¼ P,M ¼ Y RO1,RO2 ¼ ð4:57Þ
C F,RO1 CP,Pr 1  f C,RO2 þ
f C,RO2
Y RO2

If the concentrate part stream CC,RO2 from the second pass is recycled into the
feed to the first pass, the feed capacity of the pretreatment CP,Pr will be reduced by
the recycled concentrate capacity, and the recovery coefficient of the total reverse
osmosis system will increase accordingly (see Eq. 4.54). YRO1,RO2 is then calculated
according to Eq. 4.58.
252 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

C P,M Y RO1
¼ Y RO1,RO2,R ¼ ð4:58Þ
CP,Pr 1  f C,RO2 þ
f C,RO2
 ð1  Y RO1 þ Y RO1  Y RO2 Þ
Y RO2

fC,RO2 ¼ capacity factor RO2 (RO 2nd pass) [-]

As shown by both of the above equations, the recovery coefficient YRO1,RO2 of the
two-pass reverse osmosis tract is determined not only by the recovery coefficients
YRO1 and YRO2 of the interconnected RO units but also by the capacity of the second
pass, i.e. the capacity factor fC,RO2 of the post-desalination stage has an appreciable
influence on the total recovery coefficient. The degree of this influence is also
dependent on whether the concentrate from the second pass is recycled into the
feed to the first pass.
These relationships are presented in the graphs in Figs. 4.6 and 4.7.
The two graphs show how, for a constant recovery coefficient of the post-
desalination stage YRO2 of 90%, the recovery coefficient YRO1,RO2 of a two-pass
RO tract changes in response to variations in the recovery coefficient YRO1 of the
first pass between 30% and 55% and in the capacity of the post-desalination stage
between 50% and 100% of the RO system product capacity. The difference in the
two graphs lies in the fact that, in Fig. 4.6, the concentrate from the post-desalination
stage is discharged, whereas, in Fig. 4.7, it is recycled to the main desalination stage.
It is apparent that, under the presented conditions with the capacity factor of RO2
in a range of fC,RO2 ¼ 0.5–1.0, i.e. with a recovery coefficient YRO1 of the first pass of
45%, the total recovery coefficient YRO1,RO2 of the two-pass RO tract without
concentrate recycling is reduced by around 3–5% to approx. 40–42% in comparison
with the recovery coefficient YRO1 of the first pass, whereas, with concentrate
recycling, it is reduced only by around 1–2% to approx. 43–44%.

Fig. 4.6 Dependence of RO system recovery on 2nd pass recovery and capacity factor (without
concentrate recycling)
4.2 Basic Design Parameters and Conditions 253

Fig. 4.7 Dependence of RO system recovery on 2nd pass recovery and capacity factor (with
concentrate recycling)

The total recovery factor of an SWRO plant YT is calculated as the quotient of the
product capacity CP,M or nominal output CNO,SWRO and the seawater extraction
capacity or seawater feed capacity CF,SW of the plant (see Eq. 4.59 and Fig. 4.8).

C P,M C
YT ¼ ¼ NO,SWRO ð4:59Þ
CF,SW CF,SW

CF,SW ¼ extraction or feed capacity of SWRO plant [m3/d]


CNO,SWRO ¼ net output capacity of SWRO [m3/d]
YT ¼ total recovery factor of SWRO plant []

As illustrated by the block diagram of the process systems of an SWRO plant in


Fig. 4.9, calculation of its total recovery factor YT must take into consideration not
only the recovery coefficient of the reverse osmosis tract YRO1,RO2 but also the
product recovery factor of the pretreatment stage YPr. YPr is calculated from the
product capacity of the pretreatment stage and the feed capacity to the SWRO system
according to Eq. 4.60.

CP,Pr
Y Pr ¼ ð4:60Þ
C F,SW

YP,Pr ¼ product recovery factor of pretreatment stage []


254 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Fig. 4.8 Dependence of SWRO recovery on pretreatment recovery

By multiplying YRO1,RO2 by the product recovery factor YPr, the total recovery
factor YT of the SWRO plant is then obtained (see Eq. 4.61).

Y T ¼ Y RO1,RO2  Y Pr ð4:61Þ

How much pretreated product is recovered in the pretreatment stage in relation to


the amount of extracted seawater can have an appreciable influence on the total
recovery factor of the seawater desalination plant, as illustrated by the graph in
Fig. 4.8.
If, for example, under the design conditions of the second pass presented in
Fig. 4.8, the reverse osmosis systems RO1-RO2 have a recovery coefficient YRO1,
RO2 of 43%, then, if the product recovery factor YP,Pr of the pretreatment stage is
90%, the total recovery factor YT of the SWRO plant will fall to as low as around
38%.
The recovery rate of the pretreatment stage is primarily determined by the quality
of the raw water and can, for a given design of this stage, be influenced only to a
limited extent by operational measures. This applies above all to seasonally high
concentrations of suspended solids in the SWRO feed and if the pretreatment stage
uses filtration processes that are supplied directly with the seawater without prior
separation of solids. Also, additional solids can be formed in the intake to the
pretreatment stage by the dosing of coagulants/flocculants, and these solids must
be removed in addition to the naturally occurring suspended solids in the seawater.
The limited solids holding capacity of direct filtration processes results in frequent
backwashing of the filters. This high backwashing frequency increases the water
consumption of the pretreatment process and reduces its product capacity and
recovery rate, because the filtered seawater used for backwashing the filters must
previously be treated by the plant itself. Also, a lower recovery rate of the
pretreatment stage necessitates a higher seawater extraction capacity, and the
4.2 Basic Design Parameters and Conditions 255

corresponding intake systems of the SWRO plant must be capable of providing this
higher capacity.
Consequently, the recovery rate of a pretreatment stage can hardly be influenced
by modifying the mode of operation, but must be so determined by appropriate
selection and design of the pretreatment stage to minimize the impacts of variations
in seawater quality on the total product recovery factor YT of the SWRO plant. This
can be achieved by providing direct filtration stages with upstream processes for
prior separation of solids, such as sedimentation or flotation, with higher capacity for
separation of solids and lower water consumption than direct filtration processes. In
such a pretreatment stage consisting of several serially connected, different treatment
processes, the total product recovery factor YP,Pr,T of the pretreatment system is
calculated by multiplying the product recovery factors YPr,i of the individual treat-
ment processes (Eq. 4.62).

Y
n
Y Pr,T ¼ Y Pr,i ¼ Y Pr,1  Y Pr,2  . . . . . . . . . Y Pr,n ð4:62Þ
i¼1

YPr;T ¼ total product recovery factor of pretreatment []


YPr,;i ¼ product recovery factor of treatment unit of pretreatment stage []

Another possibility is to backwash the filters of the pretreatment stage with the
sufficiently available concentrate from the reverse osmosis systems. This dispenses
with the need for the water required by the pretreatment stage to come from the
process itself. However, it must be considered that, in addition to an increased salt
content, the reverse osmosis concentrate also contains high concentrations of
scalants (Ca, Sr, Ba, sulphate, and carbonate) and is oversaturated with these
substances. It must be ensured that this does not result in precipitation and scaling
in the filters that are to be backwashed. Also, this measure does not make it possible
to offset any loss of capacity of the pretreatment stage due to increased backwashing
frequency and associated production downtimes.

4.2.2 Basic Design Conditions

4.2.2.1 Overall SWRO Plant Flow and Mass Balances


An SWRO plant is made up of a number of process systems, which, in turn, consist
of a multiplicity of interconnected component groups. The block diagram in Fig. 4.9
is a schematic representation of the most important treatment systems (and their
interconnections) of a reverse osmosis seawater desalination plant for producing
drinking water.
256 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

For planning such a plant and to design its individual systems, groups, and
components, it is first necessary, on the basis of the above-described basic design
parameters, to specify and document the mass/concentration conditions between the
interconnected units at plant level, subsequently also within the treatment systems
themselves and then at group level between the individual group components. This is
done by preparing flow and mass balances both for normal plant operating
conditions and for minimum and maximum operating conditions.
The basic balances for the systems at plant level contain not only the capacity of
the main streams and part streams of the SWRO process but also the desired seawater
salt content and the seawater temperature. With these balances and with assumptions
for the various product recovery factors of the systems and the salt rejection factors
of the RO membranes, it is possible to make initial estimates for the possible
configuration of the SWRO stage and the capacity and salt content of the various
part streams. Together with the appropriate seawater analyses, these primary flow
and mass balances then form the basis for calculation of the reverse osmosis
membrane systems (see Chap. 5; Sect. 5.2). Equations that can be used in such
basic balances to calculate the part stream capacities and salt contents of the part
streams are compiled in Tables 4.6 and 4.7, respectively. The presented algorithms
can all be derived from mass balances and the basic equations for a reverse osmosis
unit according to Eqs. 4.33 to 4.38 and for a two-pass system according to Eqs. 4.48
to 4.58.

CP,inter nal

Pretreatment RO processes Potabilization


dosing dosing dosing
CB,RO2
Seawater feed cB,RO2 Output SWRO
CP,RO1
CF,SW CP,Pr CF,RO1 cP,RO1 CP,M CNO,SWRO
Pretreatment RO 1 RO 2 Potabilization
CC,RO1 CF,RO2 CP,RO2 cP,M cNO,,SWRO
cF,SW
cC,RO1 cP,RO2
CC,RO2
cC,RO2
BWW
treatment

YPr YRO1 YRO2 YT

CW,Pr CW,RO1 CW,RO2

Wastewater
treatment
Discharge SWRO to
CWT,SWRO outfall
CC,BWW Pr CD,SWRO
cWT,SWRO
CC,SWRO
cD,SWRO
cC,SWRO

Seawater/Product Waste water treated


Concentrate Chemical
Waste water Product water for internal use

Fig. 4.9 Diagram of SWRO plant flow and mass balances


Table 4.6 Capacity equations for SWRO flow and mass balance calculations
Parameters
To be
No. calculated Available Equations
SWRO part stream capacity equations
1 CF,SW CP,M; YT; YRO1,RO2; YPr C P,M C NO,SWRO
C F,SW ¼ CYP,M
T
¼ Y RO1,RO2 Y Pr ffi Y RO1,RO2 Y Pr
2 CP,Pr CP,M; CF,SW; YRO1,RO2; YPr C P,M
C P,Pr ¼ Y RO1,RO2 ¼ C F,SW  Y Pr
h  i
3a CF,RO1,R CP,M; YRO1,RO2; YRO2; fC,RO2 1 1
C F,RO1,R ¼ C P,M  Y RO1,RO2 þ f C,RO2  Y RO2 1
3b CP,Pr; CC,RO2 CF, RO1, R ¼ CP, Pr + CC, RO2
3c CF,RO1 CP,M; YRO1,RO2; CP,Pr C P,M
C F,RO1 ¼ C P,Pr ¼ Y RO1,RO2
4 CF,RO2 CP,M; fC,RO2; YRO2 C P,M  f C,RO2
C F,RO2 ¼ Y RO2
h  i
5a 1 1
4.2 Basic Design Parameters and Conditions

CC,RO1,R CP,M; YRO1; YRO1,RO2; YRO2; fC, C C,RO1,R ¼ C P,M  ð1  Y RO1 Þ  þ f C,RO2  1
Y RO1,RO2 Y RO2
RO2
5b CC,RO1 CP,M; YRO1,RO2; YRO1 C P,M
C C,RO1 ¼ Y RO1,RO2  ð1  Y RO1 Þ
 
6 CC,RO2 CP,M; fC,RO2; YRO2 1
C C,RO2 ¼ C P,M  f C,RO2  Y RO2 1
7a CC,SWRO,R CC,RO1,R CC, SWRO, R ¼ CC, RO1, R
7b CC,SWRO CC,RO1; CC,RO2 CC, SWRO ¼ CC, RO1 + CC, RO2
h  i
7c CP,M; YRO1,RO2; YRO1; YRO2; fC, RO1 1
RO1,RO2
C C,SWRO ¼ C P,M  Y1Y þ f C,RO2  Y RO2
 1
RO2
8a CW,Pr CP,M; YPr; YT C W,Pr ¼ CYP,M  ð1  Y Pr Þ
T
 
8b CP,M; YRO1,RO2; YPr C P,M
C W,Pr ¼ Y RO1,RO2  Y1Pr  1
9 CWT,SWRO CW,Pr; CW,RO1; CW,RO2 CWT, SWRO ¼ CW, Pr + CW, RO1 + CW, RO2
10 CD,SWRO CC,SWRO; CWT,SWRO CD, SWRO ¼ Cc, SWRO + CWT, SWRO
(continued)
257
Table 4.6 (continued)
258

Parameters
To be
4

No. calculated Available Equations


CF,SW ¼ seawater feed capacity to SWRO [m3/d] CC;RO1 ¼ concentrate discharge capacity RO 1st pass without concentrate recycling
CP,M ¼ RO tract product capacity [m3/d] RO2 > RO1 [m3/d]
YT ¼ total recovery factor of SWRO plant [] CC,RO2 ¼ concentrate capacity RO 2nd pass [m3/d]
YRO1 ¼ recovery coefficient RO 1st pass [] CC;RO1,R ¼ concentrate discharge capacity RO 1st pass with concentrate recycling
YRO2 ¼ recovery coefficient RO 2nd pass [] RO2 > RO1[m3/d]
YRO1,RO2 ¼ recovery coefficient RO tract (RO1 + RO2) [] CC,SWRO,R ¼ concentrate discharge capacity SWRO with concentrate recycling RO2 > RO1
YPr ¼ recovery factor of pretreatment [] [m3/d]
fC,RO2 ¼ capacity factor RO2 (RO 2nd pass) [] CC,SWRO ¼ concentrate discharge capacity SWRO without concentrate recycling
CP,Pr ¼ treated seawater feed capacity RO tract [m3/d] RO2 > RO1 [m3/d]
CF,RO1,R ¼ feed capacity to RO1 with concentrate recycling CW,Pr ¼ wastewater discharge capacity of pretreatment [m3/d]
RO2 > RO1 [m3/d] CWT,SWRO ¼ wastewater capacity SWRO [m3/d]
CF,RO1 ¼ feed capacity to RO1 [m3/d] CW, RO1 ¼ wastewater capacity RO1 [m3/d]
CF,RO2 ¼ feed capacity to RO2 [m3/d] CW, RO2 ¼ wastewater capacity RO2 [m3/d]
CD,SWRO ¼ discharge capacity of SWRO plant [m3/d]
Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .
Table 4.7 Concentration equations for SWRO flow and mass balance calculations
Parameters
To be
calculated Available Equation
SWRO part stream concentration equations
1a cF,RO1 cF,SW cF, RO1 ffi cP, Pr ffi cF, SW
cF,SW
1b cF,RO1,R cF,SW; fC,RO2; YRO1; YRO2; SPRO1; SPRO2 cF,RO1,R ¼ f
C,RO2
þ1 f C,RO2 SPRO1 ð1Y RO2 SPRO2 Þ
Y RO2
Y RO1,RO2  Y RO1  f C,RO2  Y RO2

2 cP,RO1 cF,RO1; SPRO1; RRO1 cP, RO1 ¼ cF, RO1  SPRO1 ¼ cF, RO1  (1  RRO1)
3a cC,RO1 cF,RO1; YRO1; SPRO1 RO1 SPRO1 Þ
cC,RO1 ¼ cF,RO1 ð1Y
1Y RO1
3b cC,RO1,R cF,RO1,R; YRO1; SPRO1 1Y RO1 SPRO1 Þ
cC,RO1,R ¼ cF,RO1,R ð1Y RO1

4 cF,RO2 cP,RO1 cF, RO2 ¼ cP, RO1


4.2 Basic Design Parameters and Conditions

5a cP,RO2 cF,RO2; SPRO2; RRO2 cP, RO2 ¼ cF, RO2  SPRO2 ¼ cF, RO2  (1  RRO2)
5b cF,RO1; SPRO1; SPRO2; RRO1; RRO2 cP, RO2 ¼ cF, RO1  SPRO1  SPRO2 ¼ cF, RO1  (1  RRO1)  (1  RRO2)
6a cC,RO2,R cF,RO1; YRO2; SPRO1; SPRO2 ð1Y RO2 SPRO2 Þ
cC,RO2 ¼ cF,RO1,R SPRO11Y RO2

6b cC,RO2 cF,RO1; YRO2; SPRO1; SPRO2 ð1Y RO2 SPRO2 Þ


cC,RO2 ¼ cF,RO1 SPRO11Y RO2

7a cP,M cB,RO2; cP,RO2; fC,RO2 cB, RO2 6¼ cP, RO1 6¼ cF, RO1; cP, M ¼ cB, RO2  fC, RO2  (cB, RO2  cP, RO2)
7b cF,RO1; fC,RO2; SPRO1; SPRO2 cB, RO2 ¼ cP, RO1 ¼ cF, RO1; cP, M ¼ cF, RO1  SPRO1  [1  fC, RO2  (1  SPRO2)]
7c cF,RO1; fC,RO2; RRO1; RRO2 cB, RO2 ¼ cP, RO1 ¼ cF, RO1; cP, M ¼ cF, RO1  (1  RRO1)  (1  fC, RO2  RRO2)
8a cC,SWRO Cc,RO1; Cc,RO2; cC,RO1; cC,RO2 þC C,RO2 cC,RO2
C C,SWRO ¼ C C,RO1 þ C C,RO2 ; cC,SWRO ¼ CC,RO1 cCC,RO1
C,RO1 þC C,RO2

8b cC,SWRO,R cF,RO1,R; YRO1; SPRO1 1Y RO1 SPRO1 Þ


cC,SWRO,R ¼ cC,RO1 ¼ cF,RO1,R ð1Y RO1

9 cWT,SWRO CW,Pr; CW,RO1; CW,RO2; cW,Pr; cW,RO1; cW, W,RO1 cW,RO1 þC W,RO2 cW,RO2
cWT,SWRO ¼ CW,Pr cW,PrCþC
W,Pr þC W,RO1 þC W,RO2
RO2
10 cD,SWRO CC,SWRO; CWT,SWRO; cC,SWRO; cWT,SWRO þC WT,SWRO cWT,SWRO
cD,SWRO ¼ CC,SWRO cCC,SWRO
C,SWRO þC WT,SWRO

(continued)
259
Table 4.7 (continued)
260

Parameters
To be
4

calculated Available Equation


cF,SW ¼ seawater feed concentration [mg/l] cP,M ¼ concentration of mixed final RO product [mg/l]
cP,Pr ¼ concentration of pretreatment product [mg/l] cC,SWRO ¼ concentration of RO concentrate [mg/l]
cF,RO1 ¼ concentration of feed to RO1 [mg/l] cC,SWRO,R ¼ concentration of RO concentrate [mg/l]
cF,RO1,R ¼ concentration of feed to RO1 with concentrate recycling cC,SWRO,R ¼ concentration of RO concentrate with concentrate recycling RO2 > RO1
RO2 > RO1 [mg/l] [mg/l]
cP,RO1 ¼ concentration of product of RO1 [mg/l] cWT,RO1 ¼ concentration of wastewater of RO1 [mg/l]
cC,RO1 ¼ concentration of concentrate of RO1 [mg/l] cWT,RO2 ¼ concentration of wastewater of RO2 [mg/l]
cC,RO1,R ¼ concentration of concentrate of RO1 with concentrate cWT,SWRO ¼ concentration of wastewater of SWRO [mg/l]
recycling RO2 > RO1 [mg/l] cD,SWRO ¼ concentration of discharge of SWRO [mg/l]
cF,RO2 ¼ concentration of feed to RO2 [mg/l] SPRO1 ¼ salt passage factor RO1 []
cC,RO2 ¼ concentration of concentrate of RO2 [mg/l] SPRO2 ¼ salt passage factor RO2 []
cC,RO2,R ¼ concentration of concentrate of RO2 with concentrate RRO1 ¼ salt rejection factor RO1 []
recycling RO2 > RO1 [mg/l] RRO2 ¼ salt rejection factor RO2 []
Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .
4.2 Basic Design Parameters and Conditions 261

The two tables present equations both with concentrate recycling from the second
pass (RO2) into the feed to the first pass (RO1) and without concentrate recycling.
With concentrate recycling, when mixed with the seawater feed, the lower salt content
of the concentrate recycled from RO2 into the feed to RO1 results in a lower salt
content cF,RO1,R in the seawater feed (see Table 4.7-1b) and, therefore, in a lower salt
content cC,RO1,R in the concentrate of seawater desalination stage RO1 (Table 4.7-3b).
The seawater extraction capacity of the SWRO process is reduced by the recycled
concentrate capacity from RO2 (Table 4.6-3b). All of the concentrate from the reverse
osmosis systems is discharged from seawater desalination unit RO1 (Table 4.6-7a).
Concentrate recycling from the second pass into the feed to the first pass is
standard practice in two-pass reverse osmosis systems. If, for example, the product
from the SWRO process is required to have a very low boron concentration, it may
be necessary to forgo concentrate recycling in order to ensure that the boron
accumulated in the second pass of the RO tract is not recycled back into the feed
to the first pass of the desalination process.
Figure 4.10 shows basic flow and mass balances for an SWRO plant with a
capacity of 100,000 m3/day for the standard design conditions of such a plant.
The table of balances makes a distinction between capacity and flow. The
capacity is the average of the total flows over a certain period of time. However,
the dimensioning of pipes, pump units, and process components is based on the
required range of flows. Therefore, flow and mass balances must also include the
normal, maximum, and minimum flows of the systems and components.
With the results of the membrane calculations, the basic balances are then
appropriately augmented by additional parameters, such as:

• More accurate values for the composition and total salt content of the concentrate
and product part streams
• Specification of the pressure conditions for the feed to the membranes and the
design of the necessary high-pressure pump and energy recovery systems and in
other part streams of the desalination processes
• The dosing capacities and amounts of chemical additives in the feed to the reverse
osmosis systems and the consequent changes in concentration and pH value

With the then ensuing specification of further treatment stages of the SWRO
plant, such as pretreatment, product post-treatment, and, if necessary, wastewater
treatment, the basic balance is then extended to form a comprehensive balance of the
SWRO systems. This additionally involves:

• Determination of the amounts of chemicals and their dosing rates for


pretreatment, product post-treatment, and wastewater treatment
• Preparation of balances of the wastewater quantities arising from chemical
membrane cleaning and the total amount of wastewater from the SWRO plant
• Solids balances for pretreatment, wastewater discharge, or, if necessary, waste-
water treatment, including determination of residues for disposal
• Specification of the pressure conditions in the low- and medium-pressure systems
of the SWRO plant
262 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Basic Parameter
Seawater -TDS cF,RO mg/l 40.000
Target product- TDS cP,M mg/l 200
Net product capacity CP,M m3/d 100.000
Internal consumption CP,Internal m3/d 700
Recovery factor
• Pretreatment YPr - 0,95
• RO1 YRO1 - 0,45
• RO2 YRO2 - 0,90
• RO1 + RO2 YRO1,RO2 - 0,436
• SWRO YT - 0,414
Salt rejection factor
• RO1 RRO1 - 0,99
• RO2 RRO2 - 0,97
Capacity factor RO2 fC,RO2 - 0,515

Partstream Capacity Flow TDS Partstream Capacity Flow TDS


No. m3/d m3/h mg/l No. m3/d m3/h mg/l
1 241.287 10.054 40.000 9 5.727 239 3.797
2 229.222 9.551 40.000 10 129.222 5.384 70.636
3 234.950 9.790 39.025 11 2.600 108 > 390
4 105.727 4.405 390 12 7.700 321 40.000
5 57.274 2.386 390 13 3.400 142 40.000
6 48.454 2.019 390 14 10.300 429 ~ 40,000
7 51.546 2.148 12 15 129.222 5.384 70.636
8 100.000 4.167 195 16 142.922 5.955 67.699

Fig. 4.10 Table of basic flow and mass balances of a 100,000 m3/d SWRO plant
4.2 Basic Design Parameters and Conditions 263

The amounts of chemicals DC,100% and their dosing rates RDC,100% for the various
treatment systems of the SWRO plant are calculated for each dosing point primarily
on the basis of the dosing rate RDC,100% of the relevant chemical and the flow FW into
which the chemical is to be dosed (see Eq. 4.63).

RDC,100%  F W
DC,100% ¼ ð4:63Þ
1000

RDC,100% ¼ dosing rate of chemical at 100% concentration [mg/l ¼ g/m3]


DC,100% ¼ amount of chemical at 100% concentration [kg/h]
FW ¼ flow to be dosed [m3/h]

The value of the dosing rate is normally given at 100% active concentration of the
relevant chemical RDC,100%. It is specified for the dosing points of the treatment
stages of the SWRO plant either based on experience with similar operational
systems, determined by experiment, or calculated for each dosing point from the
known or targeted compositions of the part streams of the process stage. The dosing
rate RDC,100% and the amount of each chemical DC,100% at 100% concentration are
converted to the dosing rate RDC,X% and the amount of chemical at delivery
concentration DC,X% according to Eqs. 4.64 to 4.66.

RDC,100%  100
RDC,X% ¼ ð4:64Þ
X%deliv
RDC,X%  F W
DC,X% ¼ ð4:65Þ
1000
DC,100%  100
DC,X% ¼ ð4:66Þ
X%deliv

RDC,X% ¼ dosing rate of chemical at delivery concentration of X% [mg/l ¼ g/m3]


DC,X% ¼ amount of chemical at delivery concentration of X% [kg/h]
X%deliv ¼ delivery concentration [%]

The required dosing flow FDC,X% of chemical at delivery concentration X%deliv at


its dosing point is then determined from the above-given basic parameters and the
density of the dosing solution, as shown in Eq. 4.67.

DC,X% RDC,X%  F W RDC,100%100FW


F DC,X% ¼ ¼ ¼ ð4:67Þ
ρC,X% ρC,X%  1000 X%deliv  ρC,X%  1000

FDC,X% ¼ dosing flow of chemical at delivery concentration of X% [l/h]


ρC,X% ¼ density of chemical solution at delivery concentration of X% [kg/l]
264 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Depending on the composition and/or quality of the water flow or product flow to
be treated, it is repeatedly necessary to vary, i.e. to reduce or increase, the dosing
rate. Also, the flow to be dosed is subject to variation. Both must be taken into
consideration in the dimensioning of the dosing systems, i.e. the dosing systems
must be designed to cover the entire range from the lowest to the highest dosing rate.
The results of the calculations of chemical consumption and their dosing rates as
described above are compiled in a list for all the dosing points of the SWRO plant,
not only for normal operation but also for best-case and worst-case conditions. This
list serves as the basis for including the dosing of chemicals in the flow and mass
balances of the desalination plant.
All the balances at plant level must be prepared both for the normal design
conditions of the plant and also for the minimum and maximum conditions as well
as other special operating conditions, and the resulting design parameters must be
documented in balances for the respective plant operating conditions. These mass
balances form the basis for designing the process systems of the plant. The mass
balances are usually integrated into process flow diagrams in which the process
systems of the SWRO plant are symbolically represented with their main part
streams.
In the ensuing steps of the planning process, the flow and mass balances are then
extended to the process systems level of the plant, i.e. balances are prepared and
system-specific design parameters are determined for the individual systems in the
same way as at plant level. The thus obtained results then form the basis for
designing the process systems and their component groups.
The detailed design of the SWRO plant once again requires the preparation of
flow and mass balances at group level in order to determine the parameters for the
design and technical specification of the components contained in the groups.
The flow and mass balances for the various levels of the SWRO plant, together
with their various operating conditions, form the basis for all further planning
activities, such as the preparation of detailed process flow diagrams, piping and
instrumentation diagrams (PID), instrument lists, piping isometries with pipe and
valve lists, as well as the dimensioning, detailed calculation, and technical specifica-
tion of the other mechanical equipment items of the plant.

4.2.3 Basic Planning Requirements, Their Assessment


and Definition

As described above in detail, determination, optimization, and definition of the basic


design parameters provide the basis for deciding on the configuration, design, and
dimensioning of the processes of an SWRO plant. However, the design of a desali-
nation plant is also significantly influenced by additional, external constraints and
requirements. Such factors include:
4.2 Basic Design Parameters and Conditions 265

• The scale and type of power supply to the SWRO plant


• The supply of the plant with chemicals, their availability, as-delivered condition,
and terms of delivery
• The form of integration of the desalination plant into an existing water supply
network and its mode of supply and the consequent offtake situation and
operating mode of the SWRO plant
• The storage of the product water with reference to security of supply and
operating mode of the plant
• The siting of the plant with reference to technical, environmental and socioeco-
nomic aspects and constraints
• The planning and operation of the plant with reference to ecological criteria and
environmental regulations

These in part highly complex criteria can have a significant influence not only on
the design as well as the technical and structural configuration of an SWRO plant but
also on the timeframe for realization of the project. At as early a stage as possible,
therefore, the thus arising constraints and requirements with regard to the planning/
engineering, design, and configuration of the plant should be clarified and defined,
and their impacts on process design, structural configuration, subsequent operation,
and realization horizon of the plant should be determined.

4.2.3.1 Energy Consumption and Supply Conditions

4.2.3.1.1 Energy Consumption


A reverse osmosis seawater desalination plant operates on electrical energy. The
total energy consumption of the plant EC total is made up of the energy consumption
values of its individual process and infrastructure systems according to Eq. 4.68.

E C total ¼ EC SEP þ E CPr þ E CRO þ EC Pot þ EC WWT þ E CPi þ E CAux


þ E C Inf þ E CDwp ð4:68Þ

E Ctotal ¼ energy consumption (EC) total of SWRO plant [kWh]


E CSEP ¼ EC seawater extraction, screening, and pumping systems [kWh]
E CPr ¼ EC pretreatment system [kWh]
E CRO ¼ EC reverse osmosis system with 1st and 2nd pass demand and cleaning
[kWh]
E CPot ¼ EC potabilization [kWh]
E CWWT ¼ EC wastewater treatment and sludge dewatering [kWh]
E CPi ¼ EC internal pumping necessary due to specifics in hydraulic profile of plant
[kWh]
E CAux ¼ EC auxiliaries (process air and water, process HVAC, etc.) [kWh]
E CInf ¼ EC of administration buildings, workshops and storage, laboratories,
non-process HVAC, communication systems, lighting, etc. [kWh]
EDwp ¼ EC drinking water pumping to supply network [kWh]
266 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

The energy consumption E Csystem i of the individual systems of an SWRO plant is


calculated according to Eq. 4.69 as the sum of the individual energy consumption
values EC g i of the energy consumers that are in operation at the given operating
condition (drive motors and other power-consuming systems).

X
i¼n
Ecsystem,i ¼ Ecc,g i ð4:69Þ
i¼1

E Csystem i ¼ Energy consumption of system i [kWh]


E Cg i ¼ Energy consumption of individual component, group i [kWh]

The energy consumption E C g i of a component i in turn results from its power


demand PDc,g i during the duration time τO of a certain operating cycle (Eq. 4.70).

E C g i ¼ PDc,g i  f Di,O  f P,O  τO ð4:70Þ

PDc,g i ¼ power demand of individual component, group i [kW]


fp,O ¼ factor of operational availability of plant during operation cycle []
fDi,O ¼ factor of duty of drive/consumer/system i during operation cycle ¼ 0 ! 1
[]
τO ¼ operation cycle duration [h]

However, this must take account of the operational availability fAS,O of the
systems to which the individual consumers belong during the operation cycle τO.
Calculation of the energy consumption of the individual consumer must also take
into consideration its duty time τDi,O during operation cycle tO. To calculate the
energy consumption of the individual consumer E Cg i , therefore, its power demand
PDc,g i is multiplied by a factor for the operational availability of the relevant system
and by an additional service factor fDi,O for the ratio of its duty time τDi,O to the total
operation duration of operation cycle τO (Eq. 4.71a).
The process systems have the highest energy consumption in an SWRO plant.
Their operating performance and, therefore, their operational availability are deci-
sive for the availability of the plant as a whole and the majority of its systems.
Consequently, when calculating the energy consumption, the factor for the opera-
tional availability of the respective systems fAS,O can, without risk of major inaccu-
racy, be set equal to the availability factor fP, O of the plant as a whole (see
Eq. 4.71b).

τDi,O
f Di,O ¼ ð4:71aÞ
τ0
4.2 Basic Design Parameters and Conditions 267

τS,O τ
f AS,O ¼ ffi f P,O ¼ P,O ð4:71bÞ
τ0 τ0

fAS,O ¼ factor of operational availability of system during operation cycle []


τDi,O ¼ duty time of individual component i during τ0 [h]
τS,O ¼ operation time of system during τ0 [h]
τP,O ¼ operation time of plant during τ0 [h]

The specific energy consumption SECSWRO (energy consumption per unit of


product) of an SWRO plant is calculated as the quotient of the total energy con-
sumption of the plant E Ctotal and the net output amount QNOðto Þ produced during the
selected operation cycle τO (Eq. 4.72).

E C total
SECSWRO ¼ ð4:72Þ
QNOðτ0 Þ

QNOðτ0 Þ ¼ net output amount during τ0 [m3]


SECSWRO ¼ specific energy consumption per unit of product of SWRO [kWh/m3]

The type and number of energy consumers in operation, and also their energy
consumption, vary depending on the operating condition of the desalination plant.
The number of energy consumers is subject to change especially in the case of
intermittent operation or in special operating states, such as during backwashing of
filtration systems in the pretreatment stage or during chemical membrane cleaning
and also at startup and shutdown of the plant. The reverse osmosis systems of the
plant are subject to substantial variations and peaks in energy consumption.
Depending on the temperature and salt content of the seawater as well as on the
age and fouling condition of the membranes, there is an increase in the required feed
pressure to the reverse osmosis units and, therefore, in the power consumption of the
reverse osmosis pumps. Consequently, SECSWRO is basically valid only for the
operating states of the plant during the period under consideration. This period can
be a period of normal operation of the plant, but, equally, it can contain a large
number of special operating states. The longer the selected operation cycle is for
which the specific energy consumption of the SWRO plant was measured or
calculated, the less the value for SECSWRO will be influenced by special operating
states. Therefore, values of the specific energy consumption of a reverse osmosis
desalination plant can be compared only if the period over which they were deter-
mined and the prevailing operating states and operating conditions of the plant are
known. Consequently, it is not possible to rely on values of SECSWRO to determine
the required total installed capacity of the electrical systems for supplying power to
an SWRO plant, because these will normally be average values for a given long
operation period.
268 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Determination of the power withdrawn by an SWRO plant from the power supply
network and dimensioning of the necessary electrical systems requires knowledge of
the maximum power consumption and maximum energy consumption Pc,max and
ECmax of the plant, which is determined for the operating case with the highest
number of drives and power-consuming systems in operation, all at their highest
power demand PDc,g i, max .
Another parameter of relevance for dimensioning the electrical systems of an
SWRO plant is the electric power installed in the plant. A distinction is made
between total installed operational power Po,T and total installed power PT. Both
parameters are calculated as the sum of the installed electric power ratings of the
individual motors and power consumers. Po,T is calculated by adding all the power
consumers that are in operation, whereas PT is calculated by adding all the power
consumers in the plant, i.e. including all the components that are on standby
(Eqs. 4.73 and 4.74).

X
i¼n
Po,T ¼ Pc i,O ð4:73Þ
i¼1

X
i¼n
PT ¼ Pc i ð4:74Þ
i¼1

Po,T ¼ total operational power installed [kW, MW]


PT ¼ total power installed [kW, MW]
Pc i, O ¼ power installed of operational component i [kW, MW]
Pc i ¼ power installed of component i [kW, MW]

The installed electric power Pc i of a drive or power consumer is calculated as its


maximum power demand PDc,i: max plus additional safety factors resulting from
standardization, e.g. for drive motors, and is, therefore, always higher than the
maximum power consumption of the consumer.
All these above-described parameters for power demand, energy consumption,
and installed power of the individual power consumers are combined with the
corresponding factors for their operational use under the various operating
conditions of the desalination plant in a list of motors and power consumers. Such
a list serves as a planning tool that makes it possible to summate the key data of the
individual components to obtain the corresponding parameters for the overall plant.
The lists of motors und power consumers form the basis for planning the power
supply for the individual components and systems of the SWRO plant as well as for
dimensioning the connection of the plant to the power supply network.

4.2.3.1.2 Energy Supply Conditions


An uninterruptible and demand-adapted power supply is essential for the reliable
operation and high availability of a reverse osmosis seawater desalination plant.
4.2 Basic Design Parameters and Conditions 269

Most present-day SWRO plants draw their power from conventional power supply
networks, with the power being generated in conventional fossil-fired power plants.
The power demand of some more recent large-scale SWRO plants is met by
renewable energy generated by wind farms and fed into existing power supply
networks. This type of power supply to a membrane desalination plant combines
the environmental benefits of renewable energy with the security of supply afforded
by a conventionally operated power network. Such a form of secure power supply to
an SWRO plant requires relatively little expenditure to safeguard the plant against
short-term power outages. Normally, the plant will have an oil- or gas-fuelled
emergency power generator that starts up automatically in the event of a power
outage and generates sufficient power to safely shut down the process systems and to
allow emergency operation of the plant’s infrastructure systems. As far as the reverse
osmosis systems of an SWRO plant are concerned, it is important that the emergency
power supply should ensure the displacement of seawater and concentrate as well as
the flushing of pumps, pipes, and membrane elements of the seawater desalination
stage. The same applies to the operation of valves, which must be kept in a safe
operating condition while the plant is out of operation. The plant is additionally
provided with battery systems to keep the plant’s monitoring systems operational
and to guarantee their data security.
Compared with all the other cost factors, the power costs of an SWRO plant make
up the highest proportion of the product costs. The planning and design process,
therefore, must attach particular importance to reducing the energy consumption of
the plant. This is all the more the case, the higher the specific power costs are. The
choice of energy recovery system has a significant influence on the energy consump-
tion of a membrane seawater desalination plant. Highly efficient energy recovery
systems are complex and, therefore, more expensive in terms of their equipment
costs than simpler, lower-cost systems. The latter, however, reduce the energy
consumption of the plant to a lesser extent (see also Chap. 5; Sect. 5.5). Another
possibility for reducing the energy consumption of the plant lies in the choice of
input parameters for the membrane design, e.g. lower specific loading of the
membrane elements with a consequent reduction in their operating pressure. This,
however, means an increase in the surface area, and thus in the cost, of the
membranes. Added to this, there is an increase in the product salt content, which
necessitates a higher capacity of the post-desalination stage in a two-pass plant. With
regard to plant operation, more frequent replacement and chemical flushing of the
membranes can limit the rise in pressure caused by membrane ageing and clogging.
This, however, increases the membrane costs as well as the costs of chemicals during
operation (for more information about influences on the energy consumption of an
SWRO plant, see [23], Chap. 8 in Volume 2).
To optimize the impact of power costs on the production costs of an SWRO plant,
it is necessary to study a multiplicity of aspects. Often, measures to reduce energy
consumption are accompanied by an increase in the costs of the process technology
as well as by higher operating costs of the plant. It must then be considered to what
extent the changes in costs balance each other out and what measures are needed to
minimize the production costs. Not only must the current fixed annual costs be taken
270 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

into consideration, but also life cycle calculations should be performed, ideally in
combination with sensitivity scenarios for variations in power costs over the lifetime
of the plant (see also [23], Chap. 9 in Volume 2).
In the case of stand-alone membrane desalination plants, which preferably oper-
ate on renewable energy (wind or solar) without drawing power from a power supply
network, continuous plant operation requires a bridging strategy for those periods
during which the renewable energy systems are unable to deliver sufficient power to
operate the plant. Such bridging can be provided by the usual emergency power
supply technologies used in conventional systems, such as the above-described gas-
or diesel-fuelled standby power generating sets and/or by storing surplus renewable
energy in batteries and using an inverter to convert the therein stored energy into
electricity to run the plant. Another possibility for storing energy is the use of
flywheel storage, which stores the supplied energy in the form of the rotational
energy of a rotor and which can be additionally coupled with the above-described
energy supply systems.
Such measures, which, in comparison with conventional forms of energy supply,
require both internal and external modifications to the power supply of an SWRO
plant, involve additional equipment and design costs and often also higher operating
costs of the power generation systems, all of which increase the cost of the power
supplied to and consumed by the plant. This explains why, compared with conven-
tional systems, such stand-alone membrane desalination plants, powered mainly or
exclusively by renewable energy, currently exist only in small numbers and with low
capacities.

4.2.3.2 Chemicals Supply and Storage


In an SWRO plant, many different chemicals are dosed not only into the main part
streams carrying raw water and product but also for the treatment of wastewater and
sludge in the wastewater treatment systems of the plant. Chemicals are also needed
for chemical cleaning of the membranes. The addition of chemicals for pretreatment,
desalination, and product post-treatment accounts for the majority of chemical
consumption in an SWRO plant. However, also the smaller amounts of chemicals
used for the treatment of wastewater are important for the quality of operation of the
plant. The addition of chemicals at the dosing points takes place continuously and/or
in a continuous/intermittent mode and is determined by the relevant dosing rate and
flow (see Eqs. 4.63 to 4.67). Conversely, for chemical cleaning of the membranes,
which is carried out at longer intervals, the necessary chemicals are added in batches
when the cleaning solutions are prepared.
Commercial delivery form, concentration, availability, and delivery time have a
key influence on the dimensioning and design of the chemical storage and handling
systems for bringing the chemicals to the required consistency and concentration for
dosing. These factors play a special role in determining the necessary storage time of
the relevant substances and also their type of storage and required storage space. If a
chemical is available both in solution form and as a solid product, it is necessary to
compare the costs of delivery of the liquid product with those of the solid product
plus the required solution preparation systems. A cost comparison is also necessary
4.2 Basic Design Parameters and Conditions 271

for deciding whether a chemical can be produced by means of appropriate systems


within the plant itself, such as for the production of:

• Sodium hypochlorite solution by seawater electrolysis or NaCl brine instead of


using commercially available NaOCl solution or liquid Cl2/chlorine gas
• Carbon dioxide by burning natural gas or oil in a CO2 production plant instead of
commercially available liquid CO2

The storage amount or storage volume of the chemicals to be continuously dosed


into the main part streams of an SWRO plant is usually calculated on the basis of the
daily consumption amount of each chemical based on gross capacity CGP mode of
operation of the SWRO plant ¼ DC,X%,CGP  24. This value is then multiplied by the
storage time tSt in days of storage of the relevant product. Specification of the storage
time must take into consideration the time from ordering until delivery, the delivery
time itself, the delivery interval, and a necessary safety margin for guaranteeing the
operational reliability of the plant. If the chemical is to be stored in tanks or silos, it is
additionally necessary to consider the volume of the chemical at delivery. The
required minimum storage volume is then calculated from the total volume of
chemical available for delivery and the necessary minimum chemical storage vol-
ume for bridging the ordering/delivery time.
For a liquid chemical, the storage volume StC,l,V, i.e. the necessary volume of the
storage tanks, is calculated with the density of the solution at its delivery concentra-
tion ρC,X% according to Eq. 4.75. For a solid product in powder form, the required
silo volume StC,s,V is calculated with the bulk density ρC,b as shown in Eq. 4.76,
while, for a solid product delivered in big bags, the mass storage amount StC,s,M is
calculated according to Eq. 4.77.

DC,X%,CGP  24  τSt
StC,l,V ¼ ð4:75Þ
ρC,X%

DC,X%,CGP  24  τSt
StC,s,V ¼ ð4:76Þ
ρC,b

StC,s,M ¼ DC,X%,CGP  24  t St ð4:77Þ

StC,l,V ¼ storage volume of liquid chemical at delivery concentration [m3,l]


StC,s,V ¼ storage volume of solid powder chemical at delivery concentration [m3]
StC,s,M ¼ mass storage amount of solid chemical at delivery concentration [kg]
DC,X%,CGP ¼ amount of chemical at delivery concentration of X% at gross produc-
tion capacity [kg/h]
ρC,b ¼ bulk density of solid chemical at delivery concentration of X% [kg/m3]
ρC,X% ¼ density of liquid chemical at delivery concentration of X% [kg/m3; kg/l]
τSt ¼ storage time of chemical [d]
272 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

The storage time τSt is normally between 14 and 28 days. However, in the case of
difficult-to-obtain chemicals with long and sharply fluctuating ordering and delivery
times, the storage times can be up to more than 3–4 weeks.
If the individual dosing points within the plant are supplied with chemicals from a
central store, the dosing points are equipped with holding or day tanks for local
dosing. Such tanks have a considerably smaller volume HC,l,V than the storage tanks
and are dimensioned based on the maximum dosing flow FDC,X%,max at the respec-
tive dosing point and a holding time tH (see Eq. 4.78). FDC,X%,max is also equivalent
to the maximum dosing pump capacity to be provided by the relevant dosing station
and is calculated as the quotient of the maximum dosing amount DDC,X%,max at the
respective dosing point and the density of the liquid chemical to be dosed (Eq. 4.79).

H C,l,V ¼ F DC,X%, max  τH ð4:78Þ


DDC,X%, max
F DC,X%, max = ð4:79Þ
ρC,X%

HC,l,V ¼ holding volume of chemical solution at delivery or dosing concentration


[m3,l]
DDC,X%,max ¼ maximum dosing amount of chemical at dosing or delivery concen-
tration of X% [kg/h]
FDC,X%,max ¼ maximum dosing flow of chemical at dosing or delivery concentration
of X% [l/h, m3/h]
τH ¼ Holding time of chemical solution [h]

The holding time τH is normally 24–48 h. If the solution to be dosed at the dosing
point is produced from a solid product in a dissolving station, then such a dissolving
station will consist of either two dissolving tanks, from which the chemical is dosed
alternately, or one dissolving tank and a holding tank. The holding tank can be
dimensioned according to Eq. 4.78, and the volume of the dissolving tank should, if
possible, not be smaller than that of the day tank. This is especially the case if the
dissolving process for obtaining the ready-to-dose solution takes a long time, such as
in the case of organic polyelectrolytes.
To obtain the dosing concentration of a chemical, it is often necessary for its
concentration at delivery to be reduced, i.e. its solution at delivery must be diluted.
This can be done by mixing the as-delivered solution in the holding tank with
product water. According to Eq. 4.78, however, this requires the tank to have a
larger volume, because there is an increase in the dosing flow and, therefore, in the
dosing pump capacity. Another possibility is to mix and dilute the undiluted solution
with product water in the delivery line of the dosing pump, in which case the
required amount of dilution water DDilw and flow of dilution water FDilw are
calculated according to Eqs. 4.80 and 4.81 from the dosing amount DDC,X%deliv
and dosing flow FDC,X%deliv according to the delivery concentration X%del of the
chemical and the desired dosing concentration X%dos.
4.2 Basic Design Parameters and Conditions 273

 
X%deliv
DDilw ¼ DDC,X%deliv  1 ð4:80Þ
X%dos
 
ρ X%deliv
F Dilw ¼ F DC,X%deliv  C,X%  1
ρDilw X%dos
 
X%deliv
ffi F DC,X%deliv  ρC,X%  1 ð4:81Þ
X%dos

DDC,X%deliv ¼ dosing amount of chemical at delivery concentration of X% [kg/h]


FDC,X%deliv ¼ dosing flow of chemical at delivery concentration of X% [l/h, m3/h]
DDilw ¼ amount of dilution water [kg/h]
FDilw ¼ flow of dilution water [l/h, m3/h]
ρC,X% ¼ density of liquid chemical at delivery concentration of X% [kg/l]
ρDilw ¼ density of dilution water [kg/l]
X%del ¼ delivery concentration of chemical [%]
X%dos ¼ dosing concentration of chemical [%]

The dilution water at the various dosing points of the SWRO plant is part of the
plant’s internal water demand and should be available when the plant is started or
restarted after a shutdown.

4.2.3.3 Mode of Operation and Flexibility of the Plant


Its design and configuration must allow an SWRO plant to flexibly adapt to changes
in operating conditions within the range of the specified basic operating parameters,
so that the capacity range of the plant is maintained and the required product quality
is not diminished. The extent to which, and the frequency with which, such changes
must be accommodated by the SWRO plant has a significant influence on the design
and degree of automation of the systems of the seawater desalination plant.
The following have an influence on the operating conditions of the plant:

• Variations in the temperature and salt content of the seawater


• Deterioration in seawater quality
• Variations in the required output of the plant
• Differences in the availability of power

4.2.3.3.1 Changes in Temperature and Quality of Seawater


Seasonal variations in temperature and changes in seawater salt content can be
addressed by modifying the operating pressure and product recovery rate of the
reverse osmosis systems. However, such changes will also have an influence on the
product composition, i.e. the permeate quality, especially from the first pass of the
seawater desalination stage. To allow a consistent quality of the final product from
the plant, e.g. the produced drinking water, the part stream capacity of the post-
desalination stage, the second pass of the SWRO plant, is suitably adapted so that the
274 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

mixed product conforms to the required composition of the final product. For this
purpose, it is necessary either to reduce or increase the capacity of the trains that are
in operation or to start up or shut down additional trains of the second pass.
As these changes in operating conditions are not of short-term effect, they can be
carried out manually. With a large-scale plant, however, for reasons of operational
reliability and to avoid human error, it is common practice to automate at least the
main functional sequences necessary for changing the operating conditions of the
plant.
Deterioration in seawater quality is usually attributable to an increase in the
content of suspended solids and colloidal substances due to weather influences
and/or increased biological growth. To allow the colloidal substances in seawater
to be captured, it is often necessary to increase the dosing of chemical coagulants/
flocculants, which results in the formation of additional solids. Consequently, this
imposes a heavier burden on the systems for removal of solids in the pretreatment
stage of the SWRO plant. In the case of filtration systems, this leads to shorter
filtration cycles, more frequent backwashing, and, consequently, higher in-process
water demand and wastewater volume and, therefore, to a lower product recovery in
those parts of the plant. The seawater extraction and seawater supply systems of the
SWRO plant for this case must be designed to deliver the then increased feed flow to
the pretreatment stage. For seawater of highly variable and generally inferior quality,
it may even be necessary to provide emergency systems and to keep these on standby
so that they can be put into operation if the feed conditions deteriorate significantly,
e.g. the addition of an upstream flotation stage for increased removal of solids prior
to the filtration stage.
Changes to the addition of chemicals must be made possible by suitable design of
the dosing systems. Such adjustments of the dosing flow due to quality variations in
the raw water are usually performed manually. Changes to the backwashing cycles
of filtration systems should be automated by monitoring the product water quality
(turbidity, particulate count, SDI) and the differential pressure.

4.2.3.3.2 Changes in Plant Output Capacity


Normally, a continuous mode of operation with as little change in load as possible
will give an SWRO plant the lowest operating costs, longest service life of its
systems and components, and, therefore, the lowest product water costs. In practice,
however, it is often not possible to implement such a mode of operation without
certain limitations. Therefore, the planning and design of a plant must give consid-
eration to those influences that lead to variations in load or to an intermittent
operating regime.
If the SWRO plant is required to deliver a sharply fluctuating output of product
water, it is necessary to match the plant’s product capacity to the offtake quantities.
A key criterion in this respect is the timing of the variations in product offtake as well
as the range of variation. If the required minimum and maximum capacities are
within the range of 30–120% of the plant’s normal capacity (see Sect. 4.2.1.2.3),
then the capacity can be adapted by varying the operating pressure and product
recovery rate while the plant is in operation. Such changes in operating regime are
4.2 Basic Design Parameters and Conditions 275

usually possible at short notice, i.e. in less than an hour. In the case of a large-scale
plant with many trains and a complex configuration, however, it can take several
hours. Also, in this case, if the plant is quite frequently operated with sharp variations
in production capacity, it is advisable to automate the operational sequences for
changes of capacity in order to ensure operational reliability and to minimize the
time required for such changes.
If the SWRO plant as a whole or in parts of it needs to be taken out of service for a
lengthy period of time in order to meet dispatch targets, this will significantly
diminish the plant’s capacity flexibility. This is because shutting down and starting
up the various systems of a membrane desalination plant takes a corresponding
amount of time to put the plant into a safe shutdown mode, e.g. back to a mode of
operation with appropriate product quality. During lengthy downtimes, it is neces-
sary not only to prevent corrosion and precipitation in the systems but also, and
especially, to avoid biological growth, particularly in the pretreatment stage and in
the membrane systems. The time needed, for example, for shutting down and
starting up the reverse osmosis systems, depends significantly on the duration of
the downtime as well as on the potential of the raw water to generate biological
growth.
In the case of short downtimes, e.g. around 2–7 days, biological growth in the
reverse osmosis systems can be kept within limits by means of periodic (daily)
flushing of the membranes with product water at low pressure. Depending on the size
and number of its trains, an SWRO plant can be shut down in less than an hour,
whereas this will take up to 3 or 4 h for a large-scale plant.
In the case of longer downtimes than the above-mentioned 2–7 days (depending
on the biological fouling potential of the seawater), it is necessary to conserve the
membranes with a biocide, such as bisulphite solution, after prior chemical cleaning
(see also Chap. 5; Sect. 5.5.2.5). To restart an RO train, the biocide solution must be
removed, the train must be flushed and brought up to pressure with seawater, and the
product must be discarded until the required product quality has been reached.
Compared with a short downtime with product flushing, the time required for
shutting down and restarting is correspondingly longer.
To shut down and restart all the process systems of an SWRO plant however takes
an even longer time. The time required for startup is longer than the time needed to
bring the SWRO to a complete shutdown. This is because, to start up an SWRO plant
whose individual systems have been shut down, the treatment stages must first be put
into operation one after the other, because the reverse osmosis systems cannot be
started until the pretreated water is of sufficient quality.
The same applies to product post-treatment and delivery of the final product to the
consumer. The product from the individual systems of the SWRO plant must be
discarded until the required quality standard for the downstream treatment system or
for delivery to the consumer has been reached. Depending on the treatment stage,
this so-called product drain time can take between 30 and 60 minutes (for the RO
units) or over an hour (for flocculation filtration in the pretreatment stage).
276 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Therefore, the total startup time of an SWRO plant τStart SWRO after a shutdown
consists of the startup times of the individual serially connected systems of the plant
τStart System (see Eq. 4.82).

X
n X
n
τstart SWRO ¼ τstart Seawaters þ τstart Pretri þ τstart UOi þ τstart Posttrs
i i

þ τstart Supplys ð4:82Þ

Ramping up the pressure during the restart of a membrane desalination plant must
take place over a certain time period and in a controlled manner if damage to the
membranes is to be avoided. In accordance with the specifications of membrane
manufacturers, the pressure should not be increased at a rate faster than 0.7 bar/
second, and the increase in feed flow should not exceed 5%/second of the operation
flow to the RO system (soft startup). To meet these requirements and to prevent
damage to the membranes, it is advisable to automate the startup process with regard
to control of pump capacity and valve operation.
The process of shutting down and conserving an entire SWRO plant for a lengthy
period of time (several weeks or longer), so-called mothballing, can take several
days, over a week, or even several weeks, depending on the size of the plant and its
configuration. Restarting can take a similar length of time.

4.2.3.3.3 Variations in Power Availability and Power Costs


What was described above in relation to variations in product offtake applies equally
to the operation of an SWRO plant if there are variations in power availability,
e.g. if the plant is independently operated on renewable energy without a backup
supply from a conventional power grid. As, in such a case, the power availability can
vary on a daily basis or even every few hours, an intermittent start/stop mode of
operation with product flushing, although possible, is a time-consuming process
owing to the necessary sequential startup and quality-dependent starting times of the
individual systems and is associated with higher operating costs. Also, the period of
unproductive plant capacity τd + τStart SWRO is high in relation to the period of
productive capacity τo –(τd + τStart SWRO). Therefore, to achieve a certain target gross
operating capacity CGP, the plant must operate at a correspondingly higher capacity
Cmax during its productive operating period (see Eqs. 4.83 and 4.84).

C max  ðτ0  ðτd þ τstart SWRO ÞÞ


C GP ¼ ð4:83Þ
τ0
C GP  τ0
Cmax ¼ ð4:84Þ
τ0  ðτd þ τstart SWRO Þ

CGP ¼ target gross operating capacity of plant during operating period τo [m3/h]
Cmax ¼ operating capacity during maximum capacity operating period [m3/h]
τo ¼ operation period of plant [hour]
4.2 Basic Design Parameters and Conditions 277

τd ¼ downtime period of SWRO plant [hour]


τStart SWRO ¼ starting time duration SWRO plant [hour]

In the time during which power is unavailable, it is better to maintain the plant in
operation with just a certain minimum capacity Cmin and to provide the necessary,
lower power demand from stored power. With such a mode of operation, the amount
of discarded product and the associated time for reaching the required product
quality are negligible and, consequently, the plant needs a lower product capacity
Cmax during the following maximum capacity period τ0,max (Eq. 4.85). In order to
achieve the desired target gross operating capacity CGP in this mode of operation
with a given minimum capacity Cmin provided during a given period of operation τ0,
min at minimum capacity and an operation period τ0,max with maximum production
capacity, the maximum plant operating capacity Cmax necessary during the maxi-
mum capacity operation period τ0,max calculates as shown in Eqs. 4.86 and 4.86a.

Cmin  τo, min þ C max  τo, max


C GP ¼ ð4:85Þ
τ0
CGP  τ0  C min  τo, min
Cmax ¼ ð4:86Þ
τo, max
 
τ0
Cmax ¼  1  ðC GP  C min Þ þ C GP ð4:86aÞ
τo, max

Cmin ¼ operating capacity during minimum capacity operating period τ0,min [m3/d]
τo,min ¼ duration of minimum capacity operation period [hour]
τo,max ¼ duration of maximum capacity operation period [hour]

In the case of brief or lengthy variations in power costs, the production costs of
the product water can be optimized by suitably adapting the production capacity of
the SWRO plant to the current price of power. The arid Gulf region and North Africa
are regularly subject to severe seasonal variations in the power consumption and,
therefore, in the energy generation costs of seawater desalination power plants. If
reverse osmosis seawater desalination is coupled with thermal desalination processes
in a hybrid system, then, when there is a fall in the amount of power to be exported
by the power plant, the use of membrane desalination for drinking water production
can be increased. Membrane desalination thus takes on the role of power consumer
and makes it possible to partially loosen the link between power generation and
product water production by thermal desalination at such a power plant and thereby
to improve the efficiency of the plant (see also Chap. 2, Sects. 2.4 and 2.4.2)
[17]. As, in such a case, the SWRO plant is part of an integrated supply system, it
is possible to avoid lengthy downtimes of the plant and the associated drawbacks.
Since the variations in power consumption are over a longer time period, this makes
278 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

it possible to control the timing of load changes of the reverse osmosis plant and to
limit the number of necessary stop/start cycles.
The situation is different if there are short-term, or even daily, variations in the
price of power and if the goal is to reduce their impact on water production costs by
changing the capacity of the SWRO plant. This requires a fluctuating mode of
operation with very short-term variations in plant capacity, similar to the above-
described mode of operation for the event of variations in power availability. Also, in
this case, the plant can be operated at lower capacity while the power costs are high,
with capacity then rising to the maximum when power costs are lower. However, it
must be examined whether such a mode of operation—and the associated
disadvantages with regard to the systems and components of the SWRO plant—is
actually capable of reducing the energy-related impact on the operating costs of the
plant.

4.2.3.3.4 Influence of Variations in Operation on SWRO Plant Design


The intermittent operation of an SWRO plant in alternation between periods of long
shutdown and normal load is not only time-consuming but also associated with
increased expenditure on chemicals, water consumption, energy, and manpower.
If the SWRO plant is required to deliver a certain output capacity, then, to
maintain that capacity in case of intermittent operation, the individual SWRO
systems must be designed for the maximum plant operating capacity Cmax (see
Eqs. 4.84 and 4.86). This means bigger systems than if the SWRO plant is in
continuous base load operation and, consequently, significantly higher plant costs.
If there are sharp fluctuations in the output required from the plant, then, even if
the minimum load on the SWRO plant is maintained over a long period of time, it is
considerably more advantageous to operate the SWRO plant at minimum capacity
than to completely shut it down for a long time. Especially as far as the RO units are
concerned, it must be considered that start/stop cycles will lead to changes in
pressure and flow that will subject the membrane elements to a high degree of
mechanical stress. Also, with regard to the other plant components, frequent start/
stop cycles will result in higher mechanical loading and, therefore, a shorter lifetime.
The negative impacts of load cycles and variations in operation can be mitigated
by extensive automation of the operational processes for load change, shutdown, and
startup. However, such design criteria must be included and documented in the
planning process in a timely manner. Automation requires functional descriptions of
the relevant processes, which then form the basis for the specification and detail
engineering of the supervisory control and data acquisition (SCADA) systems of the
plant.

4.2.3.4 Product Water Storage and Supply


The systems of an SWRO plant for delivery of the product water to the consumer
consist of water storage tanks and a pumping station. The pumping station is often
equipped with surge tanks to prevent or minimize pressure surges in the consumer’s
distribution network when the supply pumps are switched on and off. If the product
water is service water or drinking water, the supply systems are additionally
4.2 Basic Design Parameters and Conditions 279

provided with dosing stations for the dosing of disinfectants. The chemicals are
dosed in the feed to the storage tanks and after the supply pumps in order to prevent
contamination during water storage and in the consumer’s network.

4.2.3.4.1 Product Water Storage


The basis for dimensioning the net product water storage volume of an SWRO plant
is its net output capacity CNO and the desired storage time τStorage in hours (see
Eq. 4.87).

CNO
V Storage,N ¼  τstorage ð4:87Þ
24

CNO ¼ net output capacity [m3/d]


τStorage ¼ storage time [h]

A comparison of the net storage volumes of existing SWRO plants shows storage
times in a range between a minimum of 2 h and up to a maximum of 48 h.
This is mainly explainable by different plant sizes and operation modes, different
structures of the consumer’s supply network, available or unavailable storage
volumes in the consumer’s supply network, and different consumption
characteristics. However, the size of the required storage volume at an SWRO
plant is also affected by considerations on the part of the plant operator or the
consumer with regard to the security of supply. Not only the storage volume at the
plant is of decisive importance, but it is also necessary to provide a sufficient
redundancy of storage. For this reason, the volume is normally split into at least
two storage units, depending on the desired security of supply.
If the net storage volume of a desalination plant is split into subvolumes, with
consideration of the above-mentioned factors for determining the size of the individ-
ual subvolumes, this results in the situation presented in Eq. 4.88.

V Storage,N ¼ V Operation þ V Startup þ V Emergency þ V Firewater þ V Supply security ð4:88Þ

Operating Volume VOperation


Feed and extraction of the product water during operation of the plant take place in
the operating volume partition VOperation. If the goal is for the SWRO plant to operate
as continuously as possible with few load variations and few start/stop cycles, this
can be achieved by optimizing VOperation.
If the plant is operated with highly different minimum and maximum capacities
(see Sects. 4.2.3.3.2 and 4.2.3.3.3), then the available operating volume must be
sufficiently large to accept the additional volume that is produced during operation
of the plant at maximum capacity and which is used to bridge the time during which
the plant is shut down or operated at minimum capacity.
280 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Also, in the case of planned outages, before or after the outage, the plant can be
operated at increased capacity, according to the planned availability, in order to
make up for the capacity losses incurred during the outage (see Sect. 4.2.1.2.2). Also,
this additional volume must be taken into consideration when dimensioning the
operating volume component of the net storage volume.
Another possibility for avoiding capacity losses during planned or unplanned
outages of an SWRO plant is to design the plant with a product water storage volume
that makes it possible to buffer and offset the maximum necessary time for repair or
replacement of process-critical components as well as the associated capacity loss.
Also, in this case, the operating volume must be suitably dimensioned to accept this
additional volume of water.

Startup Volume VStartup


A certain amount of low-salinity water must always be available at the startup of an
SWRO plant to cover the internal water demand until the plant itself is able to
produce sufficient water to meet its own internal demand. This startup volume is
referred to as Vstartup in Eq. 4.88. This volume of water should always be available in
an SWRO plant to enable the plant to resume operation without problem after an
event such as an unplanned outage.

Emergency Volume VEmergency


In the case of an unplanned outage due to an emergency such as a power outage or
sudden damage event, a desalination plant should always be capable for a certain
period of time of delivering its normal, or at least a reduced, capacity. This is
necessary especially in cases where the plant is not part of an integrated water
network and is vital to the consumer’s water supply. In such a case, the net storage
volume of the plant should include a suitably dimensioned emergency volume
VEmergency that is not available for use during normal plant operation.

Firewater Volume VFirewater


Particularly if an SWRO plant is extensively independent and not connected to an
existing infrastructure from which water can be drawn, the product water storage
volume must include a suitably dimensioned firewater volume. Even if firewater can
be drawn from an existing water supply network, it is advisable, for safety reasons, to
provide an additional firewater volume at the SWRO plant. Once again, this compo-
nent of the net storage volume should always be available and should not be used for
operational purposes.

Supply Security Volume VSupply security


This volume is advisable if the consumer’s existing supply infrastructure does not
include a sufficient storage volume to cover such an event as a temporary peak water
demand. Another reason to increase the available water storage volume at the plant
may be that the offtaker/water supply utility has guaranteed its customers a very high
security of supply, which the utility wishes to meet by means of additional storage
4.2 Basic Design Parameters and Conditions 281

volume at the SWRO plant. The same applies to the plant operator if it has given an
extensive guarantee of an uninterruptible supply of product water to the consumer.

4.2.3.4.2 Product Water Supply


The dimensioning of the supply pumps, too, is based on the net output capacity CNO
of the SWRO plant. The capacity of a single supply pump Fsp,U is calculated from
CNO and the number of pumps in operation nop,n (Eq. 4.89). The capacity of the
supply pump group at the pumping station for delivery of the net output capacity to
the consumer is then calculated according to Eq. 4.90.

CNO
F sp,U ¼ ð4:89Þ
24  nop,n

F sp,G,N ¼ F sp,U  nop,n ð4:90Þ

Fsp,U ¼ design flow of supply pump unit [m3/h]


Fsp,G,N ¼ nominal flow of supply pump group [m3/h]

Normally, a redundancy of nop + 1 is chosen for such a pumping station. If the


consumer’s network is to be supplied for a certain length of time with a higher
capacity than the net output capacity, then one way of doing this is to increase the
number of standby pumps from nop + 1 to about nop + 3 or nop + 4. When additional
capacity is required, more standby pumps are put into operation. In order to maintain
a sufficient redundancy in such a case, however, there should still be one pump
available as a standby reserve, i.e. nop + 1 should be maintained. The available
additional capacity of the pumping station Fsp,G,Ov is then calculated according to
Eq. 4.91.
 
F sp,G,Ov ¼ F sp,U  nop,n þ nop,Ov ð4:91Þ

Fsp,G,Ov ¼ overload flow of supply pump group [m3/h]


nop,n ¼ number of units to be operated for nominal flow of supply pump group []
nop,Ov ¼ number of units to be operated for overload flow of supply pump group []

Alternatively, additional capacity can be made available by maintaining the


configuration of the pump group with a redundancy of nop + 1 while equipping the
individual pumps with speed control. Raising the speed, with a corresponding
increase in delivery from the individual pumps, can likewise provide the entire
pumping station with additional capacity.
282 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

4.2.3.5 Environmental and Socioeconomic Impacts and Their Influence


on the Planning and Design Process

4.2.3.5.1 Environmental and Socioeconomic Impacts


The construction and operation of a reverse osmosis seawater desalination plant can
have a number of different impacts on the location of the plant and its surroundings
(see Fig. 4.11).
On the process side, i.e. in terms of the treatment systems of the desalination plant
and during operation of the plant, these are possible environmental influences:

• On the marine and aquatic environment, i.e. seawater quality as well as marine
flora and fauna, as a result of discharge into the sea of the concentrate from the
desalination process and of the wastewater from the SWRO pretreatment stage,
which is also discharged via the outfall of the plant. The wastewater from the
chemical cleaning of the membranes of the plant is also usually mixed with the
concentrate and discharged into the sea. If a municipal sewage treatment system is
installed, then this wastewater fraction, which is not as saline as the other two

Environmental impacts

Marine & Aquatic Soil &


Air
Environment Groundwater

Marine life Water quality Contamina- Groundwater


Noise Emission
tion level

Quality of
Land use &
produced
alterations
water
SWRO plant
Impacts on Marine water
Fauna & development ,
landscape quality
flora impacts planning and Human
& nature impacts
implementaon health
Visual impacts
impact on Residues
landscape management

Occupational
Recreational
safety &
areas
health
impacts

Other Carbon
Impacts due
footprint of
Social to increased
plant
impacts traffic Climate
change
For climate impacts
Impacts on
change
cultural &
Economic planning
historic
impacts impacts
resources

Fig. 4.11 Impacts during development, planning, and implementation of an SWRO plant
4.2 Basic Design Parameters and Conditions 283

above-mentioned part streams, can be discharged into the sewer network for
treatment at the municipal sewage treatment facility.
• On the soil and groundwater as a result of storage and handling of the chemicals
required for the treatment stages of the SWRO plant. Some of these chemicals
represent a water hazard and can contaminate soil and groundwater, thereby
posing a threat to drinking water resources.

If the seawater for the SWRO plant is extracted from onshore wells or from well
galleries close to the shore, onshore groundwater can be contaminated by infiltration
with seawater or have a negative impact on the groundwater level there.

• Through residues from systems for treating the above-mentioned wastewater


from pretreatment, chemical membrane cleaning, and product water post-
treatment, sludges directly produced in the pretreatment stage of the SWRO
plant as well as replaced, spent membranes.
• Through noise and air pollution. The high-pressure pumps as well as certain
energy recovery systems of the reverse osmosis seawater desalination process
have high noise levels that necessitate the implementation of measures to mitigate
noise within the desalination plant itself and also to reduce the transmission of
noise to the surrounding area.

The direct air-side emissions of an SWRO plant are only low. Harmful vapours or
gases may be generated during filling operations or handling of some treatment
chemicals such as hydrochloric acid, sulphuric acid, chlorine, hypochlorite solution,
and ammonia (if this compound is used). However, from a global warming perspec-
tive, an SWRO plant must also be accounted for the carbon dioxide released during
the production of the energy it consumes (the plant’s carbon footprint). The same
also applies to the resulting emissions such as dust, SiO2, and NOX.
To obtain a permit for the construction and operation of an SWRO plant, these
primary process-related environmental impacts must be mitigated in conformance
with the locally valid environmental legislation to such an extent that the legally
stipulated limits and requirements are met. The necessary technical mitigation
measures have a direct influence on the design, configuration, and later operation
of the systems of an SWRO plant.
Marine and aquatic environmental impacts can be kept within limits by appropri-
ate design of the plant’s seawater extraction systems and outfall as well as by
minimizing the use of noxious oxidizing and reducing chemicals, such as chlorine
and sulphite compounds. If the seawater is extracted from onshore wells or well
galleries, appropriate hydrogeological opinions must be sought and implemented
during the design stage of the seawater extraction systems in order to guarantee
protection of the groundwater (see also [23], Chap. 4, Sect. 4.1 in Volume 2).
Wastewater from the pretreatment stage of an SWRO plant contains the solids
from the seawater in concentrated form as well as suspended solids from the dosing
of coagulants/flocculants. Especially in large-scale SWRO plants, such wastewater is
increasingly treated in wastewater treatment systems consisting of a sedimentation
284 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

stage with a dewatering system for the arising sludge. Such treatment systems are
also used to treat the wastewater from chemical membrane cleaning if discharge into
a sewer system is not possible. The treated wastewater is then mixed with the
concentrate from the desalination process for discharge into the sea through the
outfall.
The dewatered sludges arising from wastewater treatment as well as those that are
directly produced in the sedimentation or flotation stages of the pretreatment system
must then be deposited in suitable landfill sites that are protected against groundwa-
ter contamination, because the sludges are highly saline.
To protect the soil and groundwater against chemical contamination, the systems
for storing chemicals and preparing chemical solutions are equipped with appropri-
ate chemical-impermeable seals or coatings, and the transport systems are protected
against leaking. To prevent the escape of hazardous vapours or gases into the
atmosphere, the vents of chemical storage tanks are provided with absorption
vessels. If gaseous/liquid chlorine is stored, the storage room is monitored by sensors
and the escape of chlorine into the atmosphere is in many cases prevented by the
installation of an absorption system.
Noise mitigation within the reverse osmosis seawater desalination process is not
only an important element of occupational health and safety at an SWRO plant but
also serves to reduce noise pollution in the area surrounding the plant.
The carbon footprint and indirect atmospheric emissions of an SWRO plant can
be mitigated by:

• Minimizing the energy consumption of the treatment systems through the use of
high-efficiency energy recovery systems and by optimizing both the energy
consumption of the pretreatment stage and the mode of operation of the plant
(see [23], Chap. 8 in Volume 2)
• Generating the energy for the plant from renewable resources such as wind or
solar, as is already the case at some large-scale plants in Australia, where the
energy consumed by the plants is generated in wind farms and fed into existing
power supply networks
• Combining both approaches

However, the environmental impacts of an SWRO plant are not confined to the
operation of the plant. Also, its construction can have a considerable influence on the
chosen site and its environs, the environment being affected in the same ways as
described above with regard to operation of the plant, but sometimes with even
greater impact. At the marine and aquatic level, construction of the extraction and
outfall structures can have a significant effect on marine flora and fauna. The onshore
impact comes from construction noise, dust emissions, increased traffic volumes to
and from the construction site, and tipping of earth, rubble, residues, and soil as well
as groundwater pollution and wastewater discharge. Therefore, the permitting pro-
cedure includes official requirements for mitigating the environmental impacts
during construction of the plant.
4.2 Basic Design Parameters and Conditions 285

Even the early stages of the overall planning and design of the plant can help to
mitigate the environmental impacts during the construction period. This can be
achieved, for example, by the choice and design of the construction- and process-
related systems of the SWRO plant, by opting for systems that can be erected with
relatively little impact on the environment. This applies especially to the seawater
extraction systems and to the outfall from the plant as well as to the choice of
pretreatment processes [18].
Environmental influences, however, are only one aspect of the way in which an
SWRO plant impacts the site and its environs (see Fig. 4.11). Particularly if the
purpose of the plant is to produce drinking water, the quality of the water it supplies
is a key human health aspect. Here, too, the planning and design process must have
regard to national legislation at the site of the plant and/or to international guidelines
(WHO—Guidelines for Drinking-Water Quality [19], WHO—Safe Drinking-Water
from Desalination [20]) and must ensure compliance with these requirements during
operation of the SWRO plant (see also [23], Chap. 3 in Volume 2). A further human
health aspect is that of occupational health and safety during both construction and
operation of the plant. Here, too, both owner and operator must observe the relevant
national regulations and requirements. Any impacts on the technical design of the
SWRO plant must be taken into consideration in the planning and design process.
With increasing plant size, there are additional aspects that play a role in the way
an SWRO plant has an impact on the site and its environs (see Fig. 4.11). This is
especially the case where the desalination plant is a major government or local-
authority infrastructure project and is publicly financed.
An SWRO plant with a product capacity of several tens of thousands up to some
100,000 m3 per day is a large-scale infrastructure project that, with increasing plant
size, has the ability to exert a major impact on the marine and terrestrial biospheres at
the proposed site and its environs. Furthermore, projects of this scale must also take
account of aspects of regional planning as well as socioeconomic factors, such as:

• Structural and visual changes to the landscape


• Reduced availability of recreation areas and deterioration of their recreational
value
• Conflicts in relation to the preservation of existing nature reserves and historic
sites
• Negative effects on tourism in the area surrounding the site
• Increase in the volume of traffic during the construction period and increased
noise levels during construction and operation of the plant
• Fall in residential quality and price of properties close to the site
• Conflicts with existing marine structures and activities, such as commercial
fishing and fish farming, angling, sailing, and surfing
• Negative effects on existing agricultural land

The impacts of the construction and operation of an SWRO plant on the biosphere
as well as regional planning and socioeconomic aspects can have a significant
286 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

influence on the structural/architectural design of the plant and its integration into the
landscape.
On the other hand, the construction of a desalination plant can have thoroughly
advantageous consequences for the site and its environs, such as:

• A secure, high-quality water supply


• The creation of additional jobs
• A stimulus to the local and regional economy
• A rise in the incomes of local authorities close to the site
• An improvement in the standard of local and regional transport

The pros and cons of the various above-described factors in connection with the
construction and operation of a large-scale SWRO infrastructure project must be
taken into consideration in the siting of such a plant and must be weighed against
each other with due regard to their respective importance for the siting decision.

4.2.3.5.2 SWRO Plant Location, Site Search, and Selection


There are various possibilities with regard to the siting of an SWRO plant. The plant
can be:

• Integrated into an existing industrial, touristic, or municipal complex (co-siting or


co-location)
• Part of a seawater desalination power plant containing both thermal desalination
and membrane desalination systems (hybrid desalination)
• Added to a power plant (power plant co-location)
• A stand-alone plant constructed on a previously undeveloped site (greenfield
plant)

Co-siting and Co-location of SWRO Plants


As far as integration into an existing industrial, touristic, or commercial com-
plex is concerned, siting of the plant will usually be determined with respect to its
tie-in into the existing infrastructure. The requirements with regard to location and
integration of the plant into the existing structures are of a mainly technical nature.
The siting and planning of the overall plant, of which the seawater desalination is a
part, will normally also take account of the regional planning and socioeconomic
aspects described above in Sect. 4.2.3.5.1, with the consequence that it will not
usually be necessary to include those aspects in the permitting procedure for the
SWRO plant part of the overall complex.
The co-use of existing systems is often taken into consideration when planning
how to supply the plant with seawater and how to dispose of the concentrate and
wastewater. The same applies to the power supply for the SWRO plant and often
also to the storage and supply of the chemicals required for operation of the plant. In
addition, existing facilities can be co-used or suitably upgraded to allow disposal of
wastewater that cannot be discharged directly into the sea as well as disposal of
residues from the plant.
4.2 Basic Design Parameters and Conditions 287

The necessary environmental permits and associated permitting procedure are


conducted either concurrently with the permitting procedure for the complex as a
whole or as a supplement to an already existing environmental and construction
permitting process.
The permitting situation is similar if the SWRO plant is to be part of a hybrid
desalination system of a seawater desalination power plant.
In this case, at a technical level, the SWRO plant is even more highly integrated
into the processes and monitoring systems of the power plant than if it was just a
water supply facility, e.g. as part of an industrial complex. The same applies to its
supplies of power and chemicals. Co-use of the power plant’s process-related
systems not only could cover the extraction of seawater and discharge of concen-
trate, storage and supply of chemicals, and disposal of wastewater and residues but
also could include combined post-treatment of the product water from the SWRO
plant with the distillate from thermal desalination. In addition, the CO2-containing
vent gas from thermal desalination could be used to increase alkalinity (see Chap. 2,
Sect. 2.4, Fig. 2.26).
To achieve a given residual salt content in the product water from the combined
desalination processes, it is possible, by mixing the product water from SWRO with
the significantly lower-salinity distillate from the thermal process, either to entirely
dispense with the need for a second pass in the reverse osmosis system of the SWRO
plant or to design such a stage with a lower capacity.
If heated seawater is drawn from the heat reject section of the thermal desalination
process for the feed to the SWRO process, then the specific product flux of the
membranes can be improved by the higher temperature, i.e. their required operating
pressure can be reduced. In addition, using the cooling water can even out the
seasonal temperature profile of the feed to the SWRO process, thus avoiding the
need for significant pressure adjustments in the reverse osmosis system and thereby
also reducing the average annual energy consumption.
Added to this, there are other advantages of combining the membrane desalina-
tion process with the thermal desalination process in a seawater desalination power
plant, namely an increased flexibility in the operation of such a hybrid plant and,
compared with an all-thermal desalination process, a reduced internal energy con-
sumption of the desalination systems coupled to the heat and power generation of the
power plant and, therefore, higher thermal and electrical efficiencies of the dual-
purpose plant as a whole [17].
In the case of so-called co-location of an SWRO plant with a power plant, the
SWRO plant is constructed on or next to the site of the power plant to allow co-use of
the existing seawater extraction systems and outfall by both power plant and desali-
nation plant. By drawing the seawater feed for the SWRO plant from the heated
cooling water outflow from the condensers of the power plant, it is possible, as
described above for the hybrid system, to reduce the operating pressure of the
SWRO plant and, therefore, its energy consumption.
Mixing the high-salinity concentrate from the SWRO plant with the large volume
of cooling water discharged from the power plant reduces the gradient of salt content
rise at the site of discharge into the sea, which has environmental benefits.
288 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

With both co-location and hybrid systems, the proximity of the SWRO plant to
the site of power generation makes it possible for energy transmission losses in the
power network to be avoided or reduced.
In all cases where an SWRO plant is coupled to an existing facility to allow the
co-use of certain systems, this of course provides a corresponding potential to lower
both the capital expenditure and operating costs of the seawater desalination plant. In
the planning of hybrid systems and co-location systems, however, it is necessary to
give consideration to the particular requirements of membrane desalination plants in
terms of raw water quality and RO-specific modes of operation.
This applies especially to the seawater extraction systems. For power plants, these
are often in the form of open intake channels. This allows near-surface floating
impurities to enter the feed to the SWRO process, unlike in the case of deep
extraction, which is the preferred method of seawater extraction for a membrane
desalination plant. This may necessitate more extensive pretreatment of the seawater
for the SWRO process, which can reduce or even cancel out any cost savings from
co-use of the seawater extraction systems.
In cases where extraction and outfall systems already exist, it must be ensured
that, when the SWRO concentrate is additionally discharged, there is no recircula-
tion from the outfall back to the extraction system. If this happened, it could lead to
discharged impurities being recycled into the feed to the SWRO process or to an
increase in the salt content in the SWRO process. Conversely, other wastewaters
from power plant operation can have a negative impact on nearshore seawater
quality.
Also, if cooling water from the power plant is used as the feed to the SWRO
process, this can affect the operating performance of the SWRO plant. If corrosion
inhibitors and biocides are added to the feed to the condensers, it must be ensured
that these substances are compatible with the membranes and do not cause fouling.
Conversely, membrane fouling can be caused by corrosion products and heavy
metals, such as copper, nickel, and iron, contained in the cooling water.
The dosing of chlorine to prevent biofouling is carried out intermittently in
SWRO plants and at a lower dosing rate than is normally the case with once-
through seawater cooling in power plants or with thermal seawater desalination
processes. Intermittent dosing, as opposed to continuous chlorine dosing, has proved
necessary in many SWRO plants in order to contain biofouling of the membranes
(see [23], Chap. 2, Sect. 2.2.1 in Volume 2). If chlorine is continuously dosed,
especially upstream of the condensers of a power plant or in the feed to the thermal
processes in a hybrid plant, then the cooling water supplied to the SWRO process
can lead there to increased biological fouling of the RO membranes. Also, the higher
temperature of the cooling water is conducive to biological growth in the
pretreatment stage as well as in the reverse osmosis systems.
In addition, a higher operating temperature reduces the boron rejection of the
reverse osmosis membranes. To maintain a desired boron rejection rate with rising
water temperature, the pH in the feed to the reverse osmosis systems must be
increased, which means additional or increased dosing of sodium hydroxide. Partic-
ularly in tropical seas with already high water temperatures, the feed temperature of
4.2 Basic Design Parameters and Conditions 289

the cooling water to the SWRO plant must not reach or exceed the allowable
maximum temperature of around 40  C for the reverse osmosis membranes. At
certain times of the year or depending on the operating mode of the power plant, this
can rule out the use of cooling water in an SWRO plant.
The advantages and disadvantages of co-use and interconnection of power plant
and SWRO systems in hybrid or co-located plants must be carefully weighed against
each other. Thus, in cases where power plant seawater extraction systems are
co-used or where cooling water is used as the SWRO feed, the above-described
possible risks in connection with the operation of an SWRO plant have resulted in a
number of hybrid plants and co-located systems being provided with separate deep
extraction of seawater for the membrane desalination process, with the use of
cooling water in some cases being no more than an available option.

Stand-Alone (Greenfield) SWRO Plant


SWRO plants for municipal or regional drinking water supply are often of such a
size and take up such a large area that they cannot be directly integrated into existing
water supply infrastructure.
For such large-scale projects, it is then necessary to find a site that meets both the
technical and infrastructure-specific requirements of the plant, that is suitable in
relation to environmental sensitivity, and that meets with the approval of the local
population in terms of socioeconomic, regional planning, and public health aspects.
In addition to all these factors and with due regard to any required measures to
minimize the negative impacts of such an infrastructure project on the site and its
environs, it is also necessary to guarantee the profitability of the plant in terms of
both its construction and operation.
As far as the appropriate siting of such a plant is concerned, the first thing that
needs to be done is to identify various siting options and then to define the criteria by
which to compare the various possible options. Table 4.8 below assigns appropriate
evaluation criteria to the following categories:

• Technical
• Environmental
• Public health
• Socioeconomic
• Risks
• Economic

As Table 4.8 shows, the criteria are both technically qualitative and quantitative
and also monetary and non-monetary in nature. They are subdivided into main
criteria and sub-class criteria at various levels in the form of a criterion
hierarchy tree.
290 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 4.8 Site selection—evaluation criteria


Evaluation of main
Category Cai criterion Ci,L1 Evaluation of sub-class criterion Ci,Lx
Technical Site conditions Site availability/ownership
(general and Suitability for plant size and configuration of:
onshore) • Area and shape of site
• Type, shape, and size of coastal area
Distance from coastal area
Terrain conditions
Existing structures
Soil, groundwater, geological, and hydrological
conditions
Amount of seawater extraction attainable at site
Infrastructure availability and connectivity
conditions for:
• Integration into network for:
– Water supply
– Power supply
– Wastewater disposal
– Communication
• Wastewater discharge of membrane cleaning to:
– Existing sewer system
– Marine discharge with concentrate
• Access to facilities for public and goods transport
(roads, railroad)
Supply conditions for chemicals and other
consumables
Site conditions Seawater quality
(offshore) • Normal quality range
• Peak values of turbidity and pollutants
• Potential and extent of:
– Industrial discharges
– Municipal discharges
– Contamination by oil and hydrocarbons
Shore and oceanographic conditions
• Type of shore (sand, rock, etc.)
• Currents, tides, surf, etc.
• Seabed depth profile and morphology
Seawater extraction and concentrate outfall conditions
at potential extraction and outfall points and their
arrangement
• Water depth
• Flow patterns
• Locations relative to each other and distance
between
Environmental Marine life Effect of concentrate discharge on marine life
• Sensitive marine flora and fauna populations and
rare species at site
• Oceanographic conditions for concentrate
dissipation at discharge point
(continued)
4.2 Basic Design Parameters and Conditions 291

Table 4.8 (continued)


Evaluation of main
Category Cai criterion Ci,L1 Evaluation of sub-class criterion Ci,Lx
Mitigation measures necessary for:
• Amount of concentrate discharge
• Composition of pretreatment wastewater before
discharge with concentrate
Terrestrial life Effect on terrestrial life
• Sensitive terrestrial flora and fauna populations and
rare species at site
• Important area for human food production
• Protection or compensation measures necessary
Soil and Sensitive groundwater resources—soil protection
groundwater legislation
protection • Extended mitigation measures during construction
• Degree of bunding for SWRO plant chemical
storage
• Residues with high salt content—soil-protected
landfill for disposal availability
Energy Energy consumption for pumping of:
requirements • Seawater feed
• Product water export
Public health Noise and air Proximity to residential areas and landscape profile
emissions • Noise mitigation measures during construction and
plant operation
• Dust mitigation/suppression during construction
Marine water quality Importance of area for recreation and tourism
• Low
• Moderate
• High
Socioeconomic Land use Classification of area
classification of site • Undeveloped
• Industrial
• Mixed industrial/commercial/residential
• Residential
• Agricultural
• Recreational and touristic
– Moderate
– High
• Cultural/heritage area
– Moderate
– High
• Nature protected
– Moderate
– High
Compatibility with • Compatibility
adjacent land use – High
– Moderate
– Low
• Change from existing use
(continued)
292 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 4.8 (continued)


Evaluation of main
Category Cai criterion Ci,L1 Evaluation of sub-class criterion Ci,Lx
Negative impacts on • Noise
community • Visual impacts
• Traffic
• Quality of recreation
• Impact on cultural and historical resources
• Impact on protected nature resources
• Impacts on tourism
• Land value and pricing
• Impact on marine structures and activities (fishing,
etc.)
Positive impacts on • Secure and adequate water supply
community • Employment rate increase
• Local and regional business improvement
• Increase of municipality incomes
• Improvement of traffic conditions
Risks Project risks • Approval planning and licensing procedure
• Public acceptance
• Site preparation
• Site development
• Construction
• Operation
Economic Land costs • Low
• Moderate
• High
Cost of site • Preparation of site
development • Connection of site to:
– Water supply network
– Power supply
– Wastewater disposal
– Communication
– Seawater extraction facility
– Outfall facility
Cost of mitigation • Technical
measures • Environmental
• Public health
• Socioeconomic

To be able to compare these different factors and afterwards to rank the siting
options, it is necessary to apply a methodology that allows the individual criterion
parameters to be assigned to a given scale of values. In order to take account of
non-monetary aspects along with subjective stakeholder preferences, it is necessary
also to weight the evaluated criteria with regard to their significance for the interme-
diate and final results of the comparison.
A methodology that is suitable for such complex evaluations is the multi-criteria
analysis (MCA). Of the various types of MCA, the weighted sum method (WSM)
represents the best-known and simplest procedure for comparing a number of
options with their assigned criteria. Score values SCi are assigned to the individual
comparison criteria Ci and, in addition, each criterion is provided with a weight value
wCi , which characterizes the significance of the criterion within the analysis. The two
4.2 Basic Design Parameters and Conditions 293

values are multiplied according to Eq. 4.92 to obtain a weighted score value sCi for
the criterion.

s C i ¼ SC i  w C i ð4:92Þ

SCi ¼ score value for main criterion Ci


wCi ¼ weight value for main criterion Ci
sCi ¼ weighted score value for main criterion Ci

Table 4.9 presents examples of such weight and score values and their assignment
to certain evaluation criteria.
Instead of score and weight values from 1 to 10, it is, of course, possible to use
other ranges for the values. Also, the unit of measure need not be the same for
scoring and weighting.
The score values of the individual categories SCa,i for each siting option are
calculated from the total of the weighted score values of the criteria sCi,n (Eq. 4.93).
The score value for the siting option itself (SOptioni Þ is calculated from the total of the
weighted score values SCa,i of the corresponding categories (Eq. 4.94).

n¼N Ci
X
SCa,i ¼ SCi,n ð4:93Þ
n¼1

i¼N Ca,i
X
SOptioni ¼ SCa,i ð4:94Þ
i¼1

SCa,i ¼ score value of category Ca,i


SOptioni ¼ score value of option i

However, as shown in Table 4.8, the comparison criteria are often composed of a
number of sub-class criteria. This results in the above-mentioned criterion hierarchy
tree, which is then used to calculate the weighted scores from the lowest-level

Table 4.9 MCA—examples of score and weight values and assigned significance
Assigned significance
Score and weight values Scoring Weighting
1 Insufficient No importance
2 Unsatisfactory Low importance
5 Satisfactory Important
7 Good Very important
10 Favourable Extremely important
294 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

sub-class criteria up to the main criteria and to score the categories up to the overall
score for each option. The basic equation must be applied to each criterion Ci , Lx , n
at each level of the criterion hierarchy tree. The total of the weighted scores of the
n¼N
PLx
lower level sCi ,Lx ,n then serves as the score value SCi ,Lx 1,n for the relevant
n¼1
criterion at the next higher level. The basic equation (Eq. 4.92) then gives rise to the
series of equations presented in Eqs. 4.95, 4.96, 4.97, and 4.98.

sCi ,Lx ,n ¼ SCi ,Lx ,n  wCi ,Lx ,n ð4:95Þ


n¼N
XLx
sCi ,Lx ¼ sCi ,Lx ,n ¼ SCi ,Lx1 ,n ð4:96Þ
n¼1

sCi ,Lx1, n ¼ SCi ,Lx1 ,n  wCi ,Lx1 ,n ð4:97Þ


n¼NLx1
X
sCi ,Lx1 ¼ sCi, Lx1 ,n ¼ SCi ,Lx2 ,n ð4:98Þ
n¼1

SCi ,Lx ,n ¼ score value for sub-class criterion of main criterion Ci at level Lx and of
sub-class criterion number n
W Ci ,Lx ,n ¼ weight input for sub-class criterion of main criterion Ci at level Lx and of
sub-class criterion number n
wCi ,Lx ,n ¼ weight value for sub-class criterion of main criterion Ci at level Lx and of
sub-class criterion number n
sCi ,Lx ,n ¼ weighted score value for sub-class criterion of main criterion Ci at level Lx
and of sub-class criterion number n
SCi ,Lx 1,n ¼ score value for sub-class criterion of main criterion Ci at level Lx-1 and of
sub-class criterion number n
wCi ,Lx 1,n ¼ weight value for sub-class criterion of main criterion Ci at level Lx-1 and
of sub-class criterion number n
Lx ¼ level with level number x ¼ 1,2,. . . max x
n ¼ criterion number at level Lx
N Lx ¼ number of criteria at level Lx
N Ci ¼ number of main criterion Ci of category Ca,i
N Ca,i ¼ number of category Ca, i of option i

In this case, the weight value wCi ,Lx ,n for each criterion Ci, Lx, n is calculated
according to Eq. 4.99 from the quotient of the weight input W Ci ,Lx ,n of the respective
criterion and the sum of all n weight inputs of the relevant criteria at that level.
According to this calculation, the weight value wCi ,Lx ,n must then always be less than
1.
4.2 Basic Design Parameters and Conditions 295

W Ci ,Lx ,n
wCi ,Lx ,n = n¼N
ð4:99Þ
PLx
W Ci ,Lx ,n
n¼1

The weighted score value SCa,i of category Ca,i is then obtained according to
Eq. 4.100 as the sum of the weighted score values of the criteria of level 1 (sCi, L1 ,n Þ of
the criterion hierarchy tree.

n¼NCi
X
SCa,i ¼ sCi, L1 ,n ð4:100Þ
n¼1

The total score value SOptioni of option i is calculated as the sum of the score values
of its categories Ca,i, as shown in Eq. 4.94.
The calculation matrix in Annex 4.A2—Table 11 presents an example of the
sequence for such a criterion hierarchy tree calculation for determining the score
value of an option with four categories and options at four levels.
From the individual score values SOptioni of the available siting options, it is then
possible to determine the ranking of the options according to Eq. 4.101.

SOptioni
ROptioni ¼ i¼N ð4:101Þ
PO
SOptioni
i¼1

Figure 4.12 presents the sequence and individual steps of a site evaluation and
selection process using multi-criteria analysis.
After potentially suitable sites have been identified, the first thing that needs to be
done is to assign the parameters required for their evaluation to the categories
presented in Table 4.8 and then to define the relevant evaluation criteria and arrange
them in a criterion hierarchy tree (see categories and criteria in Table 4.8). Next, the
information required for determining the technical and environmental score values
of the individual criteria must be made available and compiled for each of the
candidate sites. The same applies to the technical aspects within the public health
and socioeconomics categories. To gain knowledge of the site conditions and
environmental circumstances of the individual options, it is necessary to conduct
appropriate on- and offshore studies and to evaluate the available data. With regard
to the technical parameters necessary for evaluating the suitability of the candidate
sites in terms of the required area of the plant and its connection to the existing
infrastructure, it is necessary to plan and design the plant to the stage where the
general layout, space requirements, and configuration of the individual plant systems
are known. Also, the mass balances of the plant, at least for seawater extraction and
discharge amounts, must be available.
296 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Identification of Calculation of
potential site options weighted scores for
subclass criteria,
main criteria &
categories
Definition of
evaluation main and
sub class criteria
Calculation of
weighted score for
each site option from
Development of
its categories scores
hirarchy tree of main
and sub class criteria

Calculation of ranking
Collection of of the site options
necessary
information on each
site option

Identification of
participants of
scoring team

Identification of
participants of
weighting team
Sensitivity analysis No
by changing scoring Acceptance
and weighting of ranking ?
Definition of numeric conditions and values
scoring values and
allocation to
corresponding Yes
criteria conditions

Detailed economic
comparison of
Definition of highest ranking sites
weighting values and
allocation to degree
of importance of
criteria Selection of site

Scores input to
criteria at the hirarchy
levels

Weights input to
criteria at the hirarchy
levels

Fig. 4.12 Flow diagram of site evaluation and selection process using multi-criteria analysis
4.2 Basic Design Parameters and Conditions 297

The team for defining the score values of the individual criteria should, by reason
of the multidisciplinary character of those values, be made up of experts from the
various relevant disciplines. The team can also include stakeholder representatives,
who can make suitable contributions, especially in relation to the scoring of socio-
economic aspects. The score values are dimensionless numbers to which are
assigned (see Table 4.9) certain evaluation criteria for the candidate site’s degree
of suitability for construction of the SWRO plant. In the example presented in
Table 4.9, the highest number 10 signifies the highest degree of suitability. In the
case of information based on technical and economic data, the numerical values of
the individual candidate sites should be compared and each individual site should
then be scored as a dimensionless score value in the ratio of its values to those of the
other sites. Such a procedure guarantees that all the criteria are compared on the same
scoring basis.
Appropriate explanations should be made available for each evaluation criterion
to make it clear to the members of the scoring team what score value should be
awarded for each level of site suitability.
The team for weighting the criteria and their score values can be of similar
composition to the scoring team, albeit with the inclusion of a larger number of
stakeholder representatives. The reason for this is that weighting is meant to give
preferential consideration to the influence of “soft”, non-monetary, and socioeco-
nomic aspects on site selection. It is, therefore, more subjective in nature than the
way in which the criteria are scored.
Once the teams have scored and weighted the criteria, the weighted scores are
calculated at the individual levels and for the individual criteria, after which the
weighted overall score is calculated for each option. The scores of the individual
options are then used to rank the candidate sites.
If the project participants and stakeholders are unable to agree on the ranking of
the candidate sites, it is possible, by means of a sensitivity analysis, to more closely
examine the impact of differences in scoring and weighting on site evaluation. This
involves the scoring and weighting teams making selective changes to the inputs for
certain criteria and recording the resulting impacts on the ranking. Especially as far
as “soft” criteria are concerned, this makes it possible to determine and demonstrate
how certain aspects possibly favoured by stakeholders can have a key influence on
site selection within the framework of the overall evaluation.
If the ranking of the sites is accepted by the members of the two teams, then
preferred sites should undergo an even more detailed economic evaluation with
actual cost factors, such as on the basis of the economic category in Table 4.8 and the
evaluation criteria given there. Such a cost comparison allows an appropriate
quantification of the economic impacts of each site on the construction and operating
costs of the SWRO plant. The result of this evaluation then allows a final decision to
be made on the siting of the SWRO plant.

4.2.3.5.3 Permitting Procedure: Environmental Impact Assessment


The construction and operation of an SWRO plant requires an official construction
and operation permit. Based on the national and regional environmental legislation
in force at the site of the plant, this permit lays down the requirements for the
298 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

environmentally compatible planning, construction, and operation of the plant. Such


a statutory permit for the realization of a seawater desalination project is made up of
a number of partial permits, which, according to the structure of the relevant
statutory provisions and their implementing regulations, must be granted by various
authorities at national, regional, and local levels.
The requirements under the different permitting procedures focus primarily on
measures to protect surface water, groundwater, air, and soil against environmental
impacts and to protect the site and its environs against noise from the plant.
However, depending on the size of the SWRO plant and the type of site
(co-location, co-siting, or greenfield location; see Sect. 4.2.3.5.2), the permitting
process may also give consideration to other, higher-ranking environmental aspects,
such as impacts on carbon footprint and biodiversity as well as sociological,
socioeconomic, and regional planning factors as well as issues of public health
and protection of cultural, historical, and tourism-related resources at the site (see
Fig. 4.11). Consequently, depending on the configuration, location, and size of the
seawater desalination plant, the scope and complexity of the required construction
and operation permits can vary quite considerably.
As the various environmental impacts of an SWRO plant are often subject to
different areas of environmental legislation and, therefore, to different administrative
authorities, and on account of the multidisciplinary and complex nature of the
permitting procedures, the permitting process for an SWRO plant can be extremely
wide-ranging and highly time-consuming.
An environmental impact assessment (EIA) is a process for structuring and
organizing the various technical and intangible aspects in the permit management
of infrastructure projects and other industrial, public, and municipal projects and for
progressing from a number of partial permits towards a comprehensive permitting
process.
The main aspects of an EIA are:

• A survey of the current situation and the collection of baseline data with regard to
the environmental, socioeconomic, agricultural, and cultural context of the
selected site
• The study and prediction of possible direct and indirect impacts of the project on
the baseline conditions of the site
• The identification of suitable project alternatives or minimization measures dur-
ing project execution in order to avoid, minimize, or offset such impacts
• The definition and execution of organizational and technical measures in the form
of an environmental management plan in order to monitor the specified measures
for mitigating the impacts of the project on the site and its environs and to
document the monitoring results

As the EIA is a key part of the planning process for a seawater desalination plant,
it must be so organized that any necessary changes to the construction or configura-
tion of the plant can be promptly and fully integrated into the overall planning
process.
This type of structured permitting process for identifying, minimizing, and
monitoring the environmental impacts of a project is firmly established in the
4.2 Basic Design Parameters and Conditions 299

environmental legislation of many industrialized and newly industrializing countries


as well as, to a growing extent, in the legislation of developing countries.
In the environmental legislation of, for example, the European Union and its
member states, the need for an EIA and the manner in which an EIA is conducted are
defined in Directive 2011/92/EU of the European Parliament and the European
Commission [21]. This directive has been incorporated by the EU member states
into their respective national environmental legislations, and its application is
regulated by appropriate administrative and implementing regulations. A key part
of EIA legislation is a precise definition of what type of project, and what size of
project, requires a full EIA. The EU directive lists and describes the kinds of projects
that require an EIA, sometimes with threshold values (in Annex I), and under what
circumstances the need for an EIA is to be examined on a case-by-case basis or with
threshold value targets for environmental impacts (in Annex II). With regard to
threshold value targets, the directive suggests the criteria for which such targets
should be defined (Annex III). Similar targets and requirements can be found in the
EIA legislation of other countries.
To date, seawater desalination plants have not been globally defined in EIA
environmental legislation as projects that require a mandatory EIA, such as in
Annex I of the EU directive. Unless this is explicitly regulated under the environ-
mental legislation valid for the site of the plant, therefore, the need for an EIA must
normally be examined on a case-by-case basis and/or by the definition of threshold
values, e.g. for plant capacity, or by a review of the scale of the environmental
impacts of the plant. Such so-called screening, i.e. whether and to what extent an
EIA is required, is an important part of EIA methodology.
In 2008, the United Nations Environment Programme (UNEP) published a
Resource and Guidance Handbook on how to conduct environmental impact
assessments for desalination plants [18, 22]. The Handbook presents the EIA process
techniques, contains proposals for screening and so-called scoping, i.e. the identifi-
cation and determination of the activities required for the EIA of a desalination plant,
and provides information on the structure and contents of the EIA report for a
desalination plant. In addition, it indicates the possible environmental impacts of a
desalination plant and explains how to conduct environmental monitoring.
Based on [21, 22], Fig. 4.13 presents a flow diagram of the individual steps of a
full EIA as well as the parties involved in the permitting process.
The EIA process is initiated when the proponent submits a permit application for
the construction and operation of a desalination plant to the relevant permitting
authority. The actual permit application should already be in the form of an initial
EIA and should reflect the structure and contents of an EIA report. It should
include the key basic parameters of the desalination plant, a process description,
and the plant configuration. It should also specify and describe the chosen site in
addition to giving some general information on the existing conditions at the site.
Information on the predicted environmental impacts of seawater extraction, concen-
trate/wastewater discharge, and residues disposal should also be made available in
order to allow the screening process to make the necessary assessments with regard
to the environmental impact of the plant. Information should also be provided on the
300 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Project definition

Feasibility phase and


Conceptual design

Initial Environmental 3.1


Impact Assessment Compilation of
IEIA or EI statement relevant laws,
regulations,
necessary permits
Public and responsible
Application for
involvement authorities
licence to authority

3.2
1 Project description
IEIA examination and
screening
3.3
Baseline data for
EIA required project area before
No project
Yes implementation

2
EIA scoping and TOR 3.4
preparation Impact evaluation and
analysis

3
EIA report 3.5
preparation Identification of
according to TOR mitigation measures

4 3.6
EIA report review and Update of initial plant
decision making design

EIA approved 3.7


Environmental
No
Yes mangement/
monitoring plan
5
Imposition of environmental
impact mitigation measures and
environmental monitoring & Proponent
management activities
Agency
6
Public involvement
Final update of plant
design according to
stipulations

7
Final permitting process, permits
and project license granting to
proponent

Fig. 4.13 Flow diagram of environmental impact assessment process


4.2 Basic Design Parameters and Conditions 301

anticipated size (area) of the plant as well as its layout at the proposed site. This
means that, at the time the permit application is submitted, the planning of the plant
should already be at the conceptual design stage and detailed flow and mass balances
should be available.
The screening process (step 1 in Fig. 4.13) decides whether a full EIA is required
for the permit application, whether a simplified EIA is sufficient, or whether the
necessary construction and operation permit requires several partial-permit pro-
cesses (steps 5 to 7).
If an EIA is necessary, the screening process is followed by scoping (step 2 in
Fig. 4.13), a process that determines the structure, contents, and scope of the EIA.
Terms of reference (ToR) describe in detail the information that the proponent must
make available to the relevant authority for the EIA process as well as the
investigations and studies that are required for that process and for determining the
baseline data of the selected site. The type and scope of the information to be
collected and made available under the ToR can be based, for example, on the
evaluation criteria assigned to the technical, environmental, public health, and
socioeconomic categories in Table 4.8.
On the other hand, the scoping process may define alternatives to the plant
described by the proponent in the permit application, in which case the EIA process
must compare these alternatives with the plant proposed in the permit application,
and the need for possible changes must be discussed and considered.
Another key part of the scoping process is the definition of the relevant laws,
administrative regulations, and implementing regulations as well as, resulting there-
from, the determination of the various partial permits. In addition, the scoping
process must designate the competent authority for each individual permit as well
as the higher authority that must coordinate and manage the overall EIA process.
Scoping should be carried out by a mixed team made up of independent experts
involved in the project, representatives of the relevant permitting authorities, and
public representatives, i.e. people from the section of the population affected by the
project.
The involvement of the local population in the EIA process and their participation
in its results is a key element of the permitting process. Public acceptance or the
nature and scale of public opposition in the course of the permitting process has a
quite considerable influence on the duration of the process and, therefore, on the
overall timeframe for realization of the project. Consequently, the public should be
involved as early as during the scoping process, i.e. representatives from the
population affected by the proposed site should be involved and should also have
a voice in the drafting of the ToR. Also, in the further course of the EIA process, the
public should be kept thoroughly informed about the project. Direct public partici-
pation in the EIA decision-making process can take place in step 4, i.e. during
examination of the proponent’s environmental impact assessment report and in
subsequent decision-making on mitigation measures and their monitoring during
planning, construction, and operation of the plant.
The EIA report (step 3 of the EIA process) normally includes the following key
elements and issues:
302 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

1. Methodology for performance of the EIA and preparation of the EIA report

This involves:
• Defining and explaining the structure and contents of the EIA report
• Describing the terms of reference (ToR) drafted during the scoping process, the
procedure for implementing the ToR, and the measures for acquiring all the
necessary information
• Explaining the nature, scale, and procedure for public involvement in the EIA

2. Mode of project realization and type of business structure and ownership situation
of the plant

This describes:
• The supply, construction, ownership, and operation structures of the proposed
plant (design-bid-build (DBB), design-build-operate (DBO), build-own-operate
(BOO), etc.)
• The responsibilities and accountabilities of owner, operator, or project company
in the different phases of realization and operation of the plant, including the
designation of contact persons
• The possible impacts of the implementation structures on the realization and
operation phases of the project, the permitting processes of the project, and the
environmental monitoring and environmental management of plant operation

3. Legislative and institutional basis of the EIA

This contains the results of the scoping process with regard to:
• The legislative basis for the various aspects of the EIA, such as regional, national,
and international laws, relevant administrative and implementing regulations, as
well as overarching guidelines and recommendations
• The specific basis and particular requirements of the permitting process
• The competent institutions and authorities, both with lead responsibility for the
EIA process and also for the individual partial-permit processes, including the
designation of contact persons

4. Detailed description of the proposed project

This part of the EIA report contains the detailed description of the proposed plant
and of the alternatives that were defined during the scoping process, with the:
• Goals of the project
• Description of the processes at the plant
• Information on the basic design parameters and operating parameters of the plant
as well as all environment-influencing factors
• Flow and mass balances of the plant and its systems
• Quantification and assessment of environmental impacts of the plant with regard
to scale and duration
4.2 Basic Design Parameters and Conditions 303

• Energy consumption of the plant and measures to reduce it


• Consumption of chemicals and other consumables
• Detailed information on the individual treatment systems and key components
• Timeframe for realization of the project with a breakdown of the timeframes for
planning, engineering and design, construction, erection, and commissioning
• Description of processes and measures during construction, erection,
commissioning, and various operation phases, particularly with regard to the
possible environmental impacts of such activities

5. Baseline situation of the site and its environs

This compiles all the information on the existing marine and terrestrial environmen-
tal situation at the site as well as the existing socioeconomic conditions and public
health aspects. It also describes how the information was obtained according to the
ToR of the scoping process by collection of available data and any additional studies
and investigations. Therefore, the baseline situation at the site describes the existing
situation at the site and its environs in relation to the above-described parameters. It
forms the basis for how to assess the scale and interference potential of the impacts
on the site and its environs from construction and operation of the proposed plant
and the extent to which mitigation and/or offset measures are required.

6. Impact of the proposed plant on the baseline situation at the site and its environs

This part of the EIA report describes and, as far as possible, quantifies all the
impacts of the proposed project (see Fig. 4.11 for impacts and see Table 4.8 for
evaluation criteria for technical, environmental, public health, and socioeconomic
categories). This part of the EIA report also evaluates and predicts how and on what
scale these impacts will alter the baseline situation at the site and its environs.
The same evaluations and predictions with regard to the impacts on the baseline
situation are required also for the alternatives or plant configuration options
identified in the scoping process, and the results must be compared with those of
the proposed plant.
The multi-criteria analysis (MCA), which was described in connection with the
site search for greenfield SWRO plants, can also be advantageously applied for the
comparisons and evaluations required in this step of the EIA process, especially for a
comparison of different options for the systems of the desalination plant [18].

7. Measures for mitigation, compensation, or avoidance of negative impacts

This step identifies the most suitable and economically feasible measures with
which the negative impacts of the project on the site and its environs can be
prevented or so mitigated that the remaining level of impact is acceptable to the
permitting authorities and the affected population groups. The degree of mitigation
at the legislative level is primarily determined by the relevant national, regional, or
local regulations and standards. The politically and socioeconomically required level
304 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

of mitigation or avoidance is influenced by a multiplicity of factors, particularly by


the degree to which the affected population groups are involved in the EIA process
and are provided with information about the project, the extent to which their
acceptance of the project can be secured and to which there is the possibility of
arriving at a consensus, as well as through offset measures of both a monetary and
non-monetary nature.
Insofar as changes are required to the originally submitted plant proposal as a
result of the mitigation, offset, or avoidance measures identified by the EIA process,
these changes must be explained and documented in detail in this part of the EIA
report based on the description of the plant in section 4 above.
If the EIA process proposes mitigation and avoidance measures or changes to the
design and configuration of the plant, whether or not these measures can be
incorporated at reasonable expense into the planning of the plant will depend
significantly on the time at which they are defined and initiated. The more advanced
the planning process is, or if detailed system design and component procurement has
already begun, the more expensive it will be, in terms of both cost and time, to
implement such requirements from the EIA process. Therefore, this step in the
permitting process and the definition of measures resulting from the permitting
process should take place at a time at which the overall planning process can still
be appropriately amended and at which the impact on the project timetable can be
kept within limits and remains manageable.

8. Environmental monitoring and management plan

This part of the EIA report should contain the draft environmental monitoring and
management plan. The plan should indicate:

• Which measuring and analysis methods will be used to record the environmental
impacts of the project during its construction and operation phases
• How and in what timeframe these measurements and analyses will be performed
• How the measurement and analysis results will be documented
• How environmental management will be organized and structured during the
construction and operation phases of the plant
• How and to what extent the relevant environment officers at the hierarchy levels
of environmental management will be informed about the results of environmen-
tal monitoring
• What measures will be implemented during construction and operation of the
plant and its environmental systems if targets for environmental parameters from
the official construction and operation permit are not met
• How the responsibilities and decision-making powers for the performance of such
measures are assigned within the environmental organization of the project

Similar to the environmental monitoring and management plan, this step of the
EIA process can additionally give rise to a plan for monitoring the product water
4.3 Overall Planning Process 305

quality and the relevant quality management, especially if the purpose of the plant is
to produce drinking water.
After the EIA report has been completed by the proponent and submitted to the
environmental authority with lead responsibility for the EIA process, the report is
examined by a multidisciplinary team similar to the scoping team or by the scoping
team (step 5 of the EIA process). Next, it is decided to what extent the contents of the
report are acceptable to both the permitting authorities and the stakeholders/commu-
nity for the granting of a construction and operation permit for the plant.
If the examining team is unable to reach agreement on approval of the EIA report,
then necessary changes must be defined and included in a new scoping process (step
2) and suitably amended terms of reference (ToR) must be drafted for revision of the
EIA report (repeat of step 3 of the EIA process).
If the decision is positive, the permitting authority defines and documents the
corresponding mitigation measures and the method of monitoring the environmental
situation of the plant in the form of an official final version of the environmental
monitoring and management plan (step 5).
After receiving the official EIA report, the proponent must verify whether his
plant design satisfies the authority’s requirements or whether it still needs to be
suitably adapted and modified (step 6).
In the final step of the EIA process, the lead authority and other competent
authorities prepare all the various permissions required for the construction and
operation permit for the plant. These are combined in a single permitting certificate
and presented to the proponent (step 7).

4.3 Overall Planning Process

The planning process for an SWRO plant consists of a multiplicity of planning and
design phases that take place both consecutively and concurrently. The phases are
interlinked in a variety of ways and the results of certain activities build on the inputs
from previous or simultaneously occurring design activities. Therefore, to meet the
targets for timeframes and results, it is most important to coordinate the individual
activities in such a way as to guarantee that the results from the various project
phases are made available for the next steps in the process in a timely and efficient
manner.
Figure 4.14 presents an example of the project planning phases for a greenfield
SWRO plant and the associated design phases. It also shows the key work results and
documentations that must be produced in the individual phases of the project as well
as the individual steps within those phases. The arrows in Fig. 4.14 indicate the start
of the activities and also show how far each activity extends into the project. Over
the course of the activity, its contents and results are adapted to the current level of
design detail achieved. The columns and the length of the arrows bear no relation-
ship to the practical duration of the phases and activities. The arrows that character-
ize the respective planning contents should only show the assignment of the
306 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Fig. 4.14 Project planning and engineering phases and corresponding activities
4.3 Overall Planning Process 307

activities to the respective planning and design phases and make clear their sequence
and parallelism.

4.3.1 Phases of Planning and Design and Focus of Works During


the Planning Sequences

The planning process for an SWRO plant is made up of three main phases (see
Fig. 4.14):

• The basic concept phase, which develops the holistic basis of the project and
determines the key project parameters and basic design parameters, such as the
product capacity of the plant, its capacity range, the required product quality, and
the availability of the plant (see Sects. 4.2.1 and 4.2.1.2, Table 4.1, Nos. 3–5).
• The project strategy phase, which determines the mode of operation of the SWRO
plant with reference to the various aspects of its integration into the existing water
supply systems and their demand structures as well as the necessary plant
configuration. This phase also examines and defines the redundancies for the
main components and groups of the SWRO plant and its systems; these
redundancies result from the required availability of the plant as defined in the
first planning phase (see Sect. 4.2.1.2.2). Also, additional external factors, such as
the type of energy supply as well as the availability and cost of energy, have an
influence on the structure of the plant and must be taken into consideration in the
configuration of the SWRO systems. These and other basic planning
requirements are described in Sect. 4.2.3.
The project strategy phase consists at the project design level of the following
subphases:
• Process philosophy
• Plant configuration

This planning phase includes an initial calculation of the reverse osmosis systems
in order to obtain the necessary baseline data for the flow and mass balance values
required for configuring the plant. This involves preliminary investigations and the
acquisition of information to determine the seawater composition, seawater quality,
and seasonal temperature variations (see also Sect. 4.2.1, Table 4.1, Nos. 1 and 2 and
the detailed description of activities in Sect. 4.2.1.1).
The result of the project strategy phase is a conceptual design that presents the
basic configuration of the plant and which also, as a so-called reference design, sets
the direction for the more in-depth dimensioning of the overall plant and its systems
in the form of basic parameters and key conditions/constraints. This conceptual
design is then expanded, updated, and elaborated in the following engineering phase
of the planning process. At this stage, the conceptual design also forms the basis of
the permit application for construction and operation of the SWRO plant. At the start
of the engineering phase, having been suitably expanded and elaborated, the con-
ceptual design is then integrated into the EIA process.
308 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

• The engineering phase consists at the design level of:


– Realisability phase
– Systems design
– Detail design

The realisability phase further develops and elaborates the conceptual design to
allow cost estimates of the plant systems and of the plant as a whole, with the goal of
assessing the economic feasibility of the project and the likely cost of the product
water. This phase of the design process also includes initial studies to optimize and
select cost-relevant systems and components of the SWRO plant. Selection decisions
are also made. This phase should also include the start of operation of the pilot plant,
accompanied by an in-depth determination of the seawater conditions at the site of
the plant. This, of course, assumes that, in the case of a greenfield plant, the siting of
the plant has already been finally decided. The realizability phase also sees the start
of work on the design results, which form the basis for further work on systems
design and detail design, such as the extended flow and mass balances of the overall
plant (see Sect. 4.2.2.1, Tables 4.6 and 4.7, Fig. 4.10) and its systems, detailed
general layouts of the plant, material specifications, lists of electrical consumers and
electrical single-line diagrams, calculation of key process-related systems, and
creation of their data sheets as well as conceptual construction planning and archi-
tectural design of the plant.
Systems design, often also referred to as basic design, involves a detailed
calculation of the individual plant systems. A start is made on drafting the documen-
tation that will be required in the detail design phase for selection, calculation, and
specification of the plant components and their arrangement in the engineering
drawings. The systems design phase also involves the development of more detailed
layouts of the overall plant and its systems. It additionally includes the functional
descriptions and logic diagrams of the individual operational processes of the plant
systems as a basis for designing the programmable control and monitoring systems
of the plant. Also, the piping plans are started or progressed in order to prepare the
bills of quantities and equipment lists for the detail design phase.
The detail design phase brings together the results of the systems design phase. It
also involves the complete engineering of the plant down to the construction plans as
well as the design and production drawings of the process systems. Creation of these
drawings presupposes that all the plant components have been defined and that it is
known how they will be installed. For this purpose, the detail design phase prepares
corresponding component specifications in which the required component
characteristics and dimensioning conditions are precisely described and on the
basis of which component suppliers are then able to bid and supply. The information
used for the component specifications comes from the systems design and detail
Annexes 309

design phases and should be so well-founded that it actually reflects the later
operating conditions of the components and avoids the need for costly corrective
changes caused by unreliable design information. This applies above all to the pump
systems and energy recovery systems of the reverse osmosis processes, because, if
incorrectly designed, such systems will result in an increased energy consumption of
the SWRO plant throughout its entire operating life and may also lead to mechanical
problems in the RO processes.
The construction plans and engineering drawings from the detail design phase are
appropriately updated during the construction phase of the plant and then serve as
so-called as-built drawings, which provide an important documentation basis for
operation of the plant.

Annexes

Annex 4.A1

Table 10 Seawater analysis form


Seawater analysis
Client:
Project/plant:
Type of sample:
Sampling location (for open sea sampling: distance to shore and depth):
Sampling date, time, and number:
• Date:
• Time:
• Number:
Sampling conditions (for seawater sampling: weather conditions (wind intensity, swell, etc.),
tide)
Sampling and preservation carried out by:
Laboratory (name and address)
Analysis date and number:
• Date:
• Number:
Analysis and quality control carried out by:
• Analysis
• Quality control
No. Parameter Symbol Method Unit Value
1 Seawater inorganic composition
1.1 Physical parameters
Temperature t
Conductivity at temperature t Ct
pH at temperature t pH,t
Oxidation/reduction potential at t ORPt
(continued)
310 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Table 10 (continued)
No. Parameter Symbol Method Unit Value
1.2 Ionic composition
1.2.1 Cation
• Sodium Na+
• Potassium K+
• Calcium Ca2+
• Magnesium Mg2+
• Strontium Sr2+
• Barium Ba2+
• Ammonia NH4+
1.2.2 Anion
• Chloride Cl
• Nitrate NO3
• Sulphate SO42
• Bromide Br
• Bicarbonate HCO3
• Carbonate CO32
• Fluoride F
• Borate [B(OH)4]
• Silicate (dissolved) SiO32
• Sulphide S2
1.2.3 Total dissolved solids TDS
1.2.4 Total alkalinity TA
1.2.5 Total hardness TH
1.2.6 Other components
• Boric acid B(OH)3
• Silica (colloidal) SiO2
• Carbon dioxide CO2
• Chlorine Cl2
2 Seawater quality
2.1 Particulate matter
Suspended solids content SS
Turbidity –
Particle size and distribution –
2.2 Metal foulants
Iron (total and dissolved) Fe
Manganese (total and dissolved) Mn
Alumina (total and dissolved) Al
2.3 Colloidal fouling potential
Silt density index SDI
Modified fouling index MFI0.45
2.4 Organic content sum parameter
Organic carbon content
• Total organic carbon TOC
• Dissolved organic carbon DOC
• Particulate organic carbon POC
(continued)
Annexes 311

Table 10 (continued)
No. Parameter Symbol Method Unit Value
Ultraviolet absorbance UV254
Specific UV absorbance SUVA
2.5 Synthetic organic pollutants and pollution indicators
Oil and grease –
Total petroleum hydrocarbons TPH
Surfactants anionic, non-ionic, and MBAS, CTAS,
cationic DBAS
2.6 Biological fouling potential (together with organic content sum parameter)
Biological oxygen demand BOD5
Nutrient status
• Total nitrogen Total N
• Total phosphorus Total P
• Ammonia nitrogen NH3-N
Chlorophyll A –
Pheophytin –
Algae count –
Bacterial count –
Faecal count –
3 Trace substances
3.1 Inorganic
Metals total
• Arsenic As
• Antimony Sb
• Beryllium Be
• Cadmium Cd
• Chromium Cr
• Copper Cu
• Lead Pb
• Mercury Hg
• Molybdenum Mo
• Nickel Ni
• Selenium Se
• Thallium Tl
• Uranium U
Other inorganic compounds
• Cyanide CN
• Sulphide S2
3.2 Organic
• Polycyclic aromatic hydrocarbons PAH
• Trihalomethanes THM
• Pesticides –
312 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

Annex 4.A2
Table 11 Site evaluation—mode of multi-criteria analysis (MCA) calculation
References 313

References
1. ISO International Organization for Standardization, “ISO 5667-1 Water quality -- Sampling --
Part 1: Guidance on the design of sampling programmes and sampling techniques,” ISO/TC
147/SC 6 Sampling (general methods), http://www.iso.org/iso/home/store/catalogue_tc/
catalogue_tc_browse.htm?commid¼52994, 2006.
2. ISO International Organization for Standardization, “ISO 5667-3 Water quality -- Sampling --
Part 3: Preservation and handling of water samples,” ISO/TC 147/SC 6 Sampling (general
methods), http://www.iso.org/iso/home/store/catalogue_tc/catalogue_detail.htm?csnumber¼53
569, 2012.
3. ISO International Organization for Standardization, “ISO 5667-9 Water quality -- Sampling --
Part 9: Guidance on sampling from marine waters,” ISO/TC 147/SC 6 Sampling (general
methods), http://www.iso.org/iso/home/store/catalogue_tc/catalogue_detail.htm?
csnumber¼11772, 1992.
4. Grasshoff K., Kremling K., Erhardt M., Methods of Seawater Analysis, 3rd Edition: WILEY-
VCH, 1999.
5. ASTM International, Annual Book of ASTM Standards, Section 11 - Water and Environmental
Technology, Volume 11.01 Water (I), Volume 11.02 Water (II).
6. American Public Health Association, American Water Works Association, Water Environment
Federation, Standard Methods for the Examination of Water and Wastewater, http://www.
standardmethods.org.
7. United States Environmental Protection Agency, Drinking Water Analytical Methods, http://
water.epa.gov/scitech/drinkingwater/labcert/analyticalmethods.cfm.
8. ISO International Organization for Standardization, ISO/TC 147 Water quality, http://www.iso.
org/iso/home/store/catalogue_tc/catalogue_tc_browse.htm?commid¼52834.
9. European Committee for Standardization (CEN), “CEN/TC 230 Water analysis,” http://www.
cen.eu/cen/Sectors/TechnicalCommitteesWorkshops/CENTechnicalCommittees/Pages/
default.aspx?param¼6211&title¼CEN/TC%20230.
10. DIN Deutsches Institut für Normung e.V., NA 119 Water Practice Standards Committee, NA
119 - 01-03 AA Water examination, German Standard Methods for the Examination of Water,
Waste Water and Sludge.
11. OREDA, Offshore Reliability Data Handbook, 5th ed., OREDA; www.oreda.com, 2009.
12. Procaccia, H. Aufort, P. Arsenis, S., The European Industry Reliability Data Bank EIReDA,
Crete University Press, 1998.
13. Center for Chemical Process Safety (CCPS) of the American Institute of Chemical Engineers
(AIChE), Guidelines for Process Equipment Reliability Data with Data Tables, Wiley, 1989.
14. Reliability Information Analysis Center (RIAC), Nonelectronic Parts Reliability Data,
RIAC, 2011.
15. Hajeeh M, Chaudhuri D., “Reliability and availability assessment of reverse osmosis,” Desali-
nation, vol. 130, pp. 185-192, 2000.
16. Bourouni K., “Availability assessment of a reverse osmosis plant: Comparison between Reli-
ability Block Diagram and Fault Tree Analysis Methods,” Desalination, vol. 313, pp.
66-76, 2013.
17. Ludwig H., “Hybrid systems in seawater desalination - practical design aspects, present status
and development perspectives,” Desalination, vol. 164, no. 1, pp. 1-18, 2004.
18. Lattemann S., Development of an Environmental Impact Assessment and Decision Support
System for Seawater Desalination Plants, Delft, The Netherlands: CRC Press/Balkema, 2010.
19. World Health Organization WHO, Guidelines for drinking-water quality, Fourth Edition, 2011.
314 4 Seawater Reverse Osmosis (SWRO) Plant: General System Configuration, Basic. . .

20. World Health Organization WHO, “Safe Drinking-water from Desalination,” 2011.
21. European Parliament and European Council, “Directive 2011/92/EU on the assessment of the
effects of certain public and private projects on the environment,” Official Journal of the
European Union, 28 January 2012.
22. UNEP United Nations Environment Programme, Desalination Resource and Guidance Manual
for Environmental Impact Assessments, Manama/Cairo: UNEP/ROWA Regional Office for
West Asia - WHO/EMRO Regional Office for the Eastern Mediterranean, 2008.
23. Ludwig, H., Reverse Osmosis Seawater Desalination Volume 2, Springer, 2022
Reverse Osmosis Membrane System: Core
Process of SWRO 5

Reverse osmosis membrane technology is the core process of an SWRO plant. In this
plant section, low-salinity product water is generated by desalinating seawater. The
dimensions and operating parameters of this RO section are critical for the design
and configuration of the balance of the seawater desalination plant.
The action of membrane desalination is that the solution from which the salt
content is to be removed is directed along a semipermeable membrane that selec-
tively allows the solvent through while largely retaining the solutes. By applying an
appropriate pressure to the solution side, mainly solvent is forced through the
membranes so that a proportion of the solution with a reduced salt content is
transferred to the membranes’ low-pressure side as the product, or permeate. This
is a reversal of osmosis, which is a key process of natural metabolism.
The physical principle governing osmosis is that a solution with differing saline
concentrations separated by a semipermeable membrane will strive to attain thermo-
dynamic equilibrium of the solvent’s chemical potential.
In a solution, the solvent has a lower chemical potential than would be the case if
it were to be pure because substances are dissolved in it, thereby lowering the molar
fraction of the solvent within the solution. If a saline solution is separated by a
semipermeable membrane from a similar but lower-salinity solution, the system
strives to establish thermodynamic equilibrium. Solvent from the side with the
higher chemical potential, i.e. the one that is less concentrated, naturally flows
through the membrane to the side with the lower chemical potential, that is to the
more concentrated solution (Fig. 5.1a). This continues until the chemical potential of
the solvent components is the same on both sides and the entire system is in
equilibrium, at which point solvent transport through the membranes ceases
(Fig. 5.1b).

Supplementary Information The online version contains supplementary material available at


[https://doi.org/10.1007/978-3-030-81931-6_5].

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 315
H. Ludwig, Reverse Osmosis Seawater Desalination Volume 1,
https://doi.org/10.1007/978-3-030-81931-6_5
316 5 Reverse Osmosis Membrane System: Core Process of SWRO

A B

Pressure
Pressure Osmotic pressure π

Chemical potential
solvent
Chemical potential
Molar fraction solvent
solvent
Molar fraction
solvent

Solution Solvent Solution Solvent

C
External Pressure
p>π

Pressure
Osmotic pressure π

Chemical potential
solvent

Molar fraction
solvent

Solution Solvent

Fig. 5.1 Principles of osmosis and reverse osmosis. (a) Forword osmosis. (b) Equilibrium. (c)
Reverse osmosis

This solvent flow up to attainment of thermodynamic equilibrium is termed


forward osmosis, or simply osmosis. As it proceeds, the volume on the solution
side increases, while that on the solvent side drops. The volumetric difference
between the solution and solvent sides corresponds to the osmotic pressure. The
equation for the osmotic pressure is derived from the mathematical relationship for
this thermodynamic equilibrium (Chap. 3, Sect. 3.2.3.3, Eqs. 3.111, 3.112, 3.113
and 3.114) [1].
If a pressure equal to that of the osmotic pressure is applied to the solution side,
the osmotic flow of solvent to this can be stopped. Thus, if low-salinity solvent is to
be transported from a solution through a semipermeable membrane, a pressure must
be applied on the solution side that is greater than its osmotic pressure. This is the
principle of reverse osmosis (Fig. 5.1c).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 317

5.1 Reverse Osmosis Membranes and Fundamentals of Their


Application

5.1.1 RO Membranes: Structure and Materials

The semipermeable membranes used for RO desalination on an industrial scale have


an anisotropic structure. Unlike isotropic membranes that in their configuration with
regard to both structure and material are uniform throughout, they exhibit a
non-uniform and asymmetrical, or anisotropic, structure. This membrane type is
made up of a very thin surface coating, the actual separation membranes and an
interlayer comprising layers that are first microporous and then have an increasing
pore size whose function is to impart the necessary mechanical strength and struc-
tural stability to the actual separation membrane layer and to the membrane as a
whole. The transportation and separation characteristics of such a membrane are
fixed almost entirely by the very thin separation layer on the membrane surface on
the pressure side. The porosities of the layers beneath this have to be so specified that
they minimize the resistance to product flow through the membranes. However, their
structure has to be sufficiently stable so that the operating pressure to which they are
subjected will, as far as possible, result in no mechanical changes to the active
membrane separation layer with consequent impairment of its transmission
properties.
But membranes differ not only in their physical structure, but also in the materials
used for the various membrane layers. Under this aspect, asymmetric membranes are
broken down still further into two categories, these being:

• integral-asymmetric for which all layers are of the same material and
• composite-asymmetric that are constructed of differing materials in the mem-
brane separation layer and the underlying support layers.

It was only the development of asymmetric RO membranes that made possible


the application of membrane-based desalination on an industrial scale so that fresh
water could be generated from seawater and brackish water. Compared with previ-
ous isotropic membranes, for the same salt rejection efficiency, it was possible with
this type of membrane to raise the product flux, i.e. the product flow rate per unit of
membrane area, many times over.

5.1.1.1 RO Seawater Membrane Development History


An asymmetric membrane was manufactured for the first time in 1960. It was named
the Loeb-Sourirajan membrane after the research team that developed the
manufacturing process for it at the University of California, Los Angeles (UCLA).
It consisted of partly acetylated cellulose acetate, which meant only some of the
hydroxyl groups of the cellulose structure were substituted by acetyl groups. In terms
of its structure, this membrane is of the integral-asymmetric type because both the
separation layer and the underlying support layers are of the same material, this
318 5 Reverse Osmosis Membrane System: Core Process of SWRO

being cellulose acetate. The Loeb-Sourirajan manufacturing process employs mem-


brane casting in which flat sheet membranes are generated [2, 3].
In the subsequent early 60s, this membrane and its manufacturing process were
further developed so that reverse osmosis could be applied on an industrial scale for
separation processes and desalination of salt solutions.
In 1963, a further important step in their application for reverse osmosis in
industrial-scale desalination plants to generate fresh water from brackish water and
seawater was the development of the spiral-wound process in the USA by a research
team from General Atomic with the support of the Office of Saline Water. This made
it possible to install flat membranes at a substantially higher membrane packing
density, i.e. membrane area per unit volume, within the membrane elements than had
previously been the case with the tubular modules and plate-and-frame modules used
up to then for the RO process. In 1968, the details of the construction and manufac-
ture of spiral-wound membrane elements were patented in the USA [4–6]. This
development work was the basis for the spiral-wound element technology that even
today is still overwhelmingly applied for seawater desalination (see Sects. 5.1.2
and 5.1.2.1).
In parallel with the work on flat membranes and spiral-wound technology, at the
close of the 60s integral-asymmetric membranes in the form of hollow fibres were
developed by a number of companies. The US-based Dow Chemical Company
developed a membrane in a hollow-fibre (HF) configuration based on cellulose
that had been acetylated to a greater degree—cellulose triacetate (CTA)—and in
1974, launched a hollow-fibre module onto the market. In 1978, the Japanese firm
Toyobo followed up with a hollow-fibre module likewise with fibres of cellulose
triacetate and then in 1979 with an HF module for seawater desalination. The US
firm DuPont concentrated its development activities on hollow fibres consisting of
aromatic polyamide (aramide) and, in 1974, a new desalination module, the
Permasep B-10 Permeator, was brought onto the market.
With these hollow-fibre modules, an even higher membrane packing density was
attained than had been possible with spiral-wound elements. This high packing
density, though, presented substantially greater challenges for the efficiency of
pretreatment processes in terms of particle separation and prevention of organic
and biological fouling than had been encountered up to then in conventional
processes for treating surface water.
This raised production performance per hollow-fibre module element meant that
it was now possible to apply membrane-based seawater desalination for higher plant
throughput rates. The DuPont B-10 module in particular was a trailblazer for
elevating RO membrane technology to industrial-scale plant dimensions. At its
most advanced development stage in 1992, DuPont’s B-10 polyamide hollow-fibre
module attained under standard test conditions (see Sect. 5.1.5.2.3); a production
rate exceeding 60 m3/day per permeator with a salt rejection efficiency of 99.5%.
With this performance, in the seawater desalination sector it was significantly
superior to spiral-wound modules with cellulose acetate membranes and also the
competing hollow-fibre modules with membranes of cellulose triacetate.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 319

These DuPont membranes dominated the market for seawater desalination


membranes from their introduction at the start of the 70s up to the mid-90s.
SWRO plants with production capacities of up to 90,000 m3/day were equipped
with DuPont B-10 permeators. But as well as membrane performance, it was the
quality of the user documentation and the very practical guidelines for the design and
engineering of plants using this membrane type that contributed to the success of
DuPont’s permeators.
In 1977, with his process for interfacial polycondensation, Cadotte succeeded in
developing a composite-asymmetric membrane that was eminently suitable for
seawater desalination. In this process, a very thin separation layer of cross-linked
aromatic polyamide was deposited onto an interlayer of polysulphone. In 1981, the
manufacture of these membranes was patented in the USA and this process essen-
tially still provides the basis for producing composite-asymmetric membranes, also
termed thin-film composite (TFC) membranes [7]. This membrane type, designated
FT-30, and its upgrades together with its variants that were developed particularly in
the 80s in combination with spiral-wound processes increasingly competed with the
hollow-fibre systems and also DuPont’s B-10 Permeator. A major reason for the
competitiveness of spiral-wound technology was that a large number of companies
modified and further developed composite membranes and their application in
spiral-wound elements and introduced these products to the market. Also, at a very
early stage, a standard for dimensioning spiral-wound elements was established and
all manufacturers of such membrane systems oriented themselves to this. This
specifies element dimensions of 4" (101 mm) and 8" (201 mm) for the diameter
and 40" (1016 mm) for the length. This meant that the membrane elements of the
various manufacturers were mutually interchangeable, so suppliers and operators of
membrane desalination plants were no longer tied to a specific membrane manufac-
turer. Under these competitive conditions, they could select the membrane manufac-
turer that was most favourable for their conditions, resulting in a substantial price
reduction in the market for desalination membranes.
In 1985, Dow Chemical took over the Filmtec company that had been founded by
Cadotte, ceased production of its own CTA hollow-fibre membranes and, on the
basis of the Cadotte patents and the membrane manufacturing process that it had
developed, Dow-Filmtec became one of the leading manufacturers of spiral-wound
modules and composite membranes.
Due to the increasing competitive pressure of the TFC spiral-wound modules, in
1999, DuPont decided to withdraw from the desalination membrane market, and in
2001/2002, ceased production of its Permasep-Permeators.
Most SWRO plants today operate with TFC spiral-wound polyamide elements,
although a small number of such plants—primarily located on the Red Sea and the
Arabian Gulf—are equipped with hollow-fibre membrane modules on the basis of
cellulose triacetate supplied by the membrane manufacturer Toyobo-Hollosep. In the
spiral-wound element sector, the firms formerly Dow-Filmtec, now DuPont—
Filmtec, Hydranautics and Toray have the largest market shares. The foremost
companies that manufacture membranes for seawater desalination and also brackish
320 5 Reverse Osmosis Membrane System: Core Process of SWRO

water membranes that are installed in a second pass are listed in Table 5.36 in Annex
5.A1.
Since their market introduction at the end of the 70s and beginning of the 80s,
both membrane types have undergone continuous further development with regard
to membrane materials and structure together with technical optimisation of the
elements and modules in which the membranes are installed. At the same time, the
pressure handling capability of the membrane elements has been substantially
upgraded so that they can now be loaded in some cases by up to 80 bar and even
more. In addition, the pressure loading of the membranes needed for a specified
product flux has been reduced while maintaining their salt rejection efficiency, or
even improving this.
Standard 8" spiral-wound TFC polyamide elements for seawater desalination of
diameter 7.9" (201 mm) and length 40" (1016 mm) with a salt rejection efficiency of
99.6% to 99.8% today exhibit a product output capacity of 25–50 m3/day, referred to
standard conditions.
Since its market introduction in 1975, the 8" membrane element has become
standard for large-scale membrane desalination plants. However, with the
continuing rise in production capacity of membrane desalination plants up to orders
of several 100,000 m3/day, membrane manufacturers are considering increasing this
standard size for the largest spiral-wound TFC elements to attain a higher product
performance capacity and so reduce the number of elements and modules and thus
also the capital equipment outlay for RO plants. Initial efforts in this direction were
taken in 1998 by the Graham Tek company with its 15" element for brackish water
and seawater, and in 2002, by Koch/Fluid Systems with an 18" element for treating
brackish water that was then followed by a desalination element for seawater.
In 2003/2004, a consortium consisting, among others, of the membrane
manufacturers Dow-Filmtec, Hydranautics and Toray, and backed by the U.-
S. Department of the Interior, Bureau of Reclamation, conducted an investigation
into the economic performance of spiral-wound elements with diameters of 16" and
20" for RO and nanofiltration plants in comparison with the existing 8" standard.
This found that, with 16" elements in particular, economic advantages were to be
expected, particularly for component design and mechanical equipment as well as
for the civil engineering costs of the RO section of desalination plants. As a result of
this finding, the consortium members recommended the adoption of an element
diameter of 16" (15.8"–402 mm) while retaining the length of the 8" element of 40"
(1016 mm) as the new standard for spiral-wound elements with an upgraded
throughput capacity [8].
Between 2008 and 2010, these higher capacity TFC spiral-wound modules were
placed on the market. In 2012, an industrial-scale SWRO plant was equipped with
these elements for the first time and it became operational in 2013, this being the
Sorek SWRO plant in Israel with a capacity of 624,000 m3/day. Depending on the
manufacturer and its standard test conditions (see Sect. 5.1.5.2.3), these membrane
elements have a production capacity of 100 to 150 m3/day with a salt rejection
efficiency of 99.6% to 99.8%, which amounts to a four-fold increase over the
capacity of the 8" TFC elements.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 321

In 2009, the firm NanoH2O launched onto the market for seawater desalination
membranes a new type of spiral-wound TFC polyamide module with a composite
membrane in which nanoparticles are embedded in its separation layer, termed thin-
film nanomembranes (TFN). This modification of the separation layer of TFC
membranes is based on research conducted at the University of California, Los
Angeles (UCLA), during which investigations were conducted how far the trans-
mission properties and salt rejection efficiency of TFC membranes could be
improved for brackish water desalination and for nanofiltration by embedding zeolite
nanoparticles of 50–150 nm, i.e. 0.05–0.15 μm. The UCLA research team reported
on the results in 2007 [9]. In 2012, NanoH2O and several employees in the USA
named in the patent specification patented a process for manufacturing separation
membranes by means of interfacial polymerization combined with embedding of
nanomaterials [10]. Following a number of tests conducted under seawater desalina-
tion conditions as encountered in practice [11], starting in 2011 the first SWRO
plants were equipped with 8" spiral-wound elements with TFN membranes.
On the basis of nanomembrane technology and its further development, also with
the embedment of other types of nanomaterials in the separation layer of TFC
membranes, impetus could be given to seawater desalination in terms of raising
product flux and so also reducing membrane operating pressure, thus cutting the
energy consumption of SWRO plants. This applies also for salt rejection efficiency
and the extent of membrane compacting, i.e. long-term membrane behaviour.
The availability of membrane elements with high unit capacities is critical for
economic performance when constructing and operating large-scale seawater mem-
brane desalination plants with production capacities that are sufficient for supplying
towns, regions, and industrial complexes with fresh water and drinking water. But
also for smaller communities, hotels, tourist centres, and companies, today seawater
desalination has advanced to become a key component of their water supply
infrastructure. RO membrane desalination has likewise established itself for com-
mercial shipping as well as for sport and leisure boating activities. For such SWRO
plants with the smaller capacities that they require, membrane manufacturers also
carry in their product portfolios a broad range of membrane elements with reduced
performance parameters, for example, 2" and 4" spiral-wound elements.

5.1.1.2 Membrane Structure and Material


Figure 5.2 shows a schematic cross-section of a composite thin-film polyamide
composite membrane and an integral-asymmetric cellulose acetate membrane stating
the approximate thickness of each membrane layer.
According to the membrane manufacturers, the active separation layer that
determines the characteristics of the RO membranes, i.e. their product flux and salt
rejection efficiency, is 0.1–0.5 μm thick, depending on membrane type. TFC
membranes are usually constructed with three layers, these being the semipermeable
separation layer (or functional barrier layer), the microporous interlayer, and the
macroporous reinforcing and supporting fabric, although they may include addi-
tional layers. For TFC membranes, the thickness of the interlayer and the reinforcing
layer backing the separation layer is 140–180 μm, while for cellulose triacetate
322 5 Reverse Osmosis Membrane System: Core Process of SWRO

Composite thin-film polyamide membrane


Nanomembrane with zeolite particles

Thickness of layer [μm]


Semipermeable membrane -
~ 0.1 – 0.2 Functional barrier layer –
Material: Aromatic polyamide
Microporous interlayer –
~ 40 - 60
Material: Polysulfone

~ 100 - 120 Reinforcing and supporting


fabric– Material: Polyester

Integral – asymmetric cellulose acetate membrane


Semipermeable membrane -
~ 0.1 – 0.5 Functional barrier layer –
Material: Cellulose triacetate

~ 80 - 100 Microporous support matrix–


Material: Cellulose triacetate

Fig. 5.2 Schematic cross-sections through asymmetric RO membranes

membranes it is around 80–100 μm. If the CTA membrane is executed as a flat sheet
membrane, an additional fabric layer may be included for reinforcement, so the
thickness of the entire membrane at about 0.2 mm will then be almost equal to that of
a TFC membrane.
For TFC membranes that find application as RO desalination membranes, the
separation layer material is an aromatic polyamide whose particular type depends on
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 323

Fig. 5.3 RO membranes: approximate chemical structure of active membrane separation layer

the manufacturer. As material for the underlying microporous interlayers,


polysulphones are used with further fabric layers usually of various types of
polyester.
Integral-asymmetric CTA membranes are wholly of cellulose triacetate, i.e. both
separation layer and reinforcing layers are of this material. Figure 5.3 shows the
approximate chemical structure of the separation layer of a TFC membrane1 of
aromatic polyamide and that of a CTA membrane.2
The transmission properties and saline retention efficiencies of an RO membrane
are determined primarily by the polymer material selected for its separation layer.
The material selected for the separation layer, however, greatly influences its surface
characteristics with regard to both physical and chemical properties. These

1
DuPont/Dow Filmtec, FT30—Membrane.
2
Toyobo, Hollosep—Membrane.
324 5 Reverse Osmosis Membrane System: Core Process of SWRO

characteristics, in turn, fix the membrane’s fouling behaviour and moreover have an


additional impact on salt transport through it.
Such surface characteristics of separation membranes are:

• roughness
• hydrophilic or hydrophobic properties
• polarity and magnitude of their electric charge.

The membrane’s fouling behaviour may be influenced by all the above


characteristics of the separation layer. The rougher the surface, the more particles
will be deposited on it. If the surface is primarily hydrophilic, preferentially water
molecules will be attracted to it and form an aqueous layer that then reduces the
adhesion of solids to the membrane’s surface. The extent to which the electric charge
favours or hinders fouling of the membrane surface depends on the status of the
charge of the inorganic or organic particles that come into contact with the separation
layer.
Investigations specifically of the organic fouling and biofouling of RO
membranes during seawater desalination have shown that separation membranes
with low roughness and hydrophilic characteristics whose surface is negatively
charged or uncharged have the least tendency to become fouled.
A surface charge, though, can influence salt transport through a semipermeable
separation membrane because an excessive concentration of oppositely charged ions
can build up at the membrane’s surface. Thus, a charge potential—the Donnan
potential—is created between this stratum and the membrane that influences salt
transport through the latter and thus, for certain ions, their degree of rejection. A
further factor is that such a charge potential is a function of pH with regard to its
nature and magnitude, and it builds up differently depending on whether the solution
is in the acidic, neutral or alkaline range. This also helps to explain why salt rejection
in RO membranes is highly pH-dependent for specific ions.
By modifying the physical and chemical structures as well as the make-up of the
surface of separation membranes during polymerization or also through appropriate
post-treatment, the fouling behaviour and salt rejection efficiency of RO membranes
can be influenced in a targeted approach [12]. In this way, for example, low-fouling
(LF) RO membranes can be manufactured.
Investigations at TFN membranes that have been generated by interfacial poly-
merization and at the same time encapsulation of zeolithic nanoparticles within the
separation layer have revealed that, compared with standard TFC membranes with
the same polymer composition of the separation layer, they exhibit significantly
different surface properties and permeabilities. Strongly hydrophilic characteristics
and an increase in the negative surface charge were observed. Also, the permeability
of the separation membranes increased for the same salt rejection capacity [9]. From
this, it may be concluded that, through further development of nanomembranes, a
corresponding potential for reducing their fouling tendency with a further improve-
ment of the transmission properties of RO desalination membranes may be expected.
Also, with regard to membrane compaction, it would appear that there is still scope
for improvement by embedding nanoparticles in TFC membranes.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 325

5.1.2 RO Membrane Elements and Modules in Application in SWRO

5.1.2.1 Spiral-Wound Membrane Elements


Thin-film composite membranes in the form of TFC/TFN flat sheet membranes find
application in industrial and large-scale membrane desalination plants as spiral-
wound elements.
For manufacturing these elements, a porous permeate collection spacer is placed
between two flat sheet membranes in a back-to-back configuration, i.e. with their
respective active separation membrane layers on the outsides, and the edges of this
membrane sandwich are bonded together on three sides. This combination of two
separation membranes and permeate collection layer is termed a membrane leaf, but
is also referred to as a membrane envelope. The only still open sides of all of the
membrane envelopes of the spiral-wound element are connected by a perforated
tube, this being the central permeate tube of the element. Permeate that passes
through the two separation membranes of the membrane envelope is directed via
the permeate collection layer to the central permeate tube. Here the permeate flow
from all membrane envelopes is brought together and directed through this header
tube to the end of the spiral-wound element.
Mesh screens are inserted between the membrane envelopes, these being the feed
spacers. Because this layer between the membrane envelopes is open at its ends, the
solution to be purified can be fed into the spiral-wound element and the concentrate
can flow out. These feed spacers provide the channels for feed flow into the element
and to the surfaces of the separation membranes and for outflow of the concentrate.
They provide the means for establishing defined flow conditions. When
manufacturing a spiral-wound element, depending on the element diameter, a
defined number of membrane envelopes together with their associated feed spacers
are wound around the central tube, so forming the cylindrical spiral-wound mem-
brane element (Fig. 5.4). The membrane winding is then covered on the outside with
a protective sleeve of fibreglass fabric.
The configuration of the flow channels formed by the feed spacer is critical for the
performance of the spiral-wound elements. This pertains to their dimensioning,
i.e. the distance created by the spacer between the separation membranes of the
membrane envelopes, as well as for the mesh structure and the material it is made
of. The dimensions of the feed channels and the type and structure of the feed spacers
determine the concentration polarization and thus the salt rejection efficiency, the
membrane area and thus production performance, the pressure loss, and the
elements’ fouling behaviour.
The thickness of the feed spacer fabric of the spiral-wound elements is between
0.71 mm (28 mil3) and 0.86 mm (34 mil). Most spiral-wound elements have a spacer
thickness of 0.71 mm. However, as the thickness of the spacer decreases, the
differential pressure of the element increases and it is more prone to fouling. Larger
spacer thicknesses improve the fouling characteristics, which means that low fouling

3
1 mil ¼ 0.001" ¼ 0.0254 mm.
326 5 Reverse Osmosis Membrane System: Core Process of SWRO

Brine Seal Endcaps

Perforated Permeate Central Tube Feed

Feed Spacer
Membrane
Feed Flow
Permeate
(through feed spacer)

Permeate Collecon Material


Concentrate
Membrane
Feed Spacer
Element Covering Layer

Fig. 5.4 Spiral-wound membrane element

Snap Ring End Cap Brine Seal Pressure Vessel Perforated Permeate Central Tube

Permeate
Feed

Concentrate

Interconnector

RO membrane elements

Fig. 5.5 RO module with spiral-wound membrane elements

elements usually have a feed spacer thickness of 0.86 mm with a reduced element
membrane area.
The membrane elements are fitted with plastic caps at each end whose geometry
differs depending on membrane manufacturer. Their purpose is to distribute the
element’s inflow and outflow uniformly over its cross-section.
Under standard test conditions and depending on the testing parameters of the
membrane manufacturer (see Sect. 5.1.5.2.3), an RO spiral-wound element attains a
product recovery rate of 7–10%. In order to obtain units with higher recoveries and
production performance as required for industrial membrane desalination plants,
multiple spiral-wound elements are mounted in a pressure vessel and connected in
series. Usually six or seven but also up to eight elements are so connected. This
results in a total length of the resulting RO modules of ~6700 mm if six elements are
installed and up to ~8800 mm with eight elements.
The solution is fed into the desalination module at its head end and flows into the
first element where salt elimination commences, as described above (Fig. 5.5). The
concentrate generated in the first element flows into the second element and subse-
quently into the other downstream spiral-wound elements. In doing so, the salt
content of the solution fed into each element and the concentrate there generated
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 327

continually increase with the number of elements. The permeate generated in each
element enters the module’s central perforated permeate tube and then exits at its
discharge end. Here also the concentrate exits the RO module (Fig. 5.5) or it exits the
module at both ends, if, for example, the module’s spiral-wound elements are
operated in a split-flow configuration.
Infeed and concentrate discharge of the RO module are by way of ports in its base
and sealing plates at its head and discharge ends. This manner of infeed and
concentrate flow is referred to as “end-port design” of the RO module and its
configuration is as shown schematically in Fig. 5.5. In the side port design of the
pressure vessel, it is provided at its infeed and discharge sides with lateral ports in its
wall for feeding the solution into the module and discharging the concentrate from
it. Like for the end-port design, permeate flows out centrally at the head and/or
discharge ends of the module.
Each spiral-wound element is fitted with a seal at its head end caps, termed a brine
seal, to seal off the annular space between elements and pressure vessel wall so that
no liquid can laterally bypass the elements.
The elements are normally joined to each other by interconnectors so that the
central permeate tubes of the spiral-wound elements form a common central perme-
ate collection tube for the module. These interconnectors between the elements are
highly critical for ensuring reliable operation and desalination performance of the
resulting module. The quality of the permeate exiting the RO module greatly
depends on the integrity of this seal as any leaks at one or more of the interconnectors
from the pressurized module sections would allow the ingress of saline concentrate
into the central low-pressure permeate tube. This could severely impact the quality
of the product from the RO module.
The connections between the spiral-wound elements in the pressure vessel of an
RO module must meet the following requirements:

• uniform distribution of infeed over the element cross-sections


• leaktightness and mechanical integrity of the connection
• prevention or minimization of telescoping
• low-pressure loss over the connection sections
• ease of handling when assembling the pressure vessel and removal of the
elements from it.

The usual means of joining the RO elements is by interconnectors in the form of


tubular sections with O-rings at each end that are inserted into the permeate tube of
each element. One RO membrane manufacturer meets the above criteria without
using such interconnectors just by means of a special design of the end caps of his
RO elements.4 Figure 5.6 shows a comparison of the two types of connection for the
RO membrane elements.
Loading of the pressure vessel of an RO osmosis module with 800 spiral-wound
elements is shown in Fig. 5.7.

4
iLEC technology of DuPont/DOW Filmtec.
328 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.6 Spiral-wound elements in the RO module: comparison between conventional connection
and iLEC technology

Fig. 5.7 Loading of 8" RO pressure vessel (Source: Author)

In their transportation condition with liquid filling, the elements weigh 15–20 kg;
so, with this weight, their manual handling and loading into the pressure vessel
present no problem.
However, this is not the case for the 16" RO membrane elements. Figure 5.8
shows a visual comparison of these two types of element. The manufacturers state a
weight of 65–75 kg for these large RO elements when they are filled with liquid, so
their transportation and handling as well as their loading into the pressure vessels are
only possible with mechanical aids.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 329

Fig. 5.8 Comparison of 8" and 16" element dimensions (Source: Author)

Fig. 5.9 Loading of a 16" element pressure vessel (Courtesy: DuPont/DOW Filmtec)

This means that hoists are needed for transportation to the place of loading and,
for the actual loading operation, special equipment is required for positioning the
element in readiness for loading, balancing it, inserting it into the pressure vessel,
and there fixing it in place. Figure 5.9 shows such loading equipment with a 16" RO
element ready for insertion into the open pressure vessel.
330 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.1.2.2 Hollow-Fibre Membrane Systems


Hollow-fibre cellulose acetate membranes as they find application for seawater
desalination are made up of fine fibres with an outside diameter of 140 μm
(0.14 mm) and a bore of 55 μm. The active membrane separation layer is located
on the external fibre surface. Thus, the fibre is pressurized by the feed flow on its
external surface and the resulting permeate collects in its internal capillary bore.
From there, it can exit at both fibre ends (Fig. 5.10).
In a hollow-fibre (HF) RO membrane element, the hollow fibres are bundled with
their ends embedded in two epoxy resin fibre sheets. Depending on the configuration
of the RO membrane element, the fibre sheets are so arranged that, at one of the two
sheets, the fibre capillary bores are closed off and permeate can only exit at the other
end of the bundle. Alternatively, the two fibre sheets are left open, so that permeate
can be discharged from the fibre bundle at both ends. The fibres are bundled in a
special cross-wound configuration that fixes them in place and ensures uniform
throughflow. Figure 5.11 shows an RO membrane element of the manufacturer of
this cellulose triacetate HF system, Toyobo Hollosep, by means of which the
permeate is drawn from both ends of the membrane bundle, referred to as a both
open-ended (BOE) element.
The membrane bundle is arranged around a central concentric double-walled
tube that serves both for admitting the feed solution to be desalinated to the
bundle and for drawing off the permeate. The bundle is fed via the outer annulus
of the double-walled tube and from there the saline solution is directed radially
through the bundle to its periphery. The concentrate that is so generated then
flows out from the bundle to between the outside of the membrane element and
the wall of the pressure vessel in which the element is installed. The permeate
product that emerges from the open fibre ends into the fibre sheets is collected in

CTA Hollow fiber

Outer diameter
140 μm
Feed

Inner diameter
55 μm

Permeate

Permeate

Active separation membrane layer

Feed

Fig. 5.10 CTA hollow-fibre membrane: dimensions and flows


5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 331

Fig. 5.11 Hollow-fibre RO membrane element (Courtesy: Toyobo-Hollosep)

permeate chambers located between the fibre sheets and the supporting plates at
each end of the element. From there, the permeate is directed into the permeate
tube in the central double-walled tube of the bundle for discharge from the
element.
With hollow-fibre membrane technology, the membrane elements are also
mounted in pressure vessels, but the membrane packaging density of HF elements
is substantially greater than that of spiral-wound elements. Under its standard test
conditions (see Sect. 5.1.5.2.3), the membrane manufacturer, Toyobo Hollosep,
specifies a product recovery of 30% for the HF elements. This means that, for
seawater desalination and depending on the seawater’s salt content and the
operating conditions of the RO system, one or two of these elements per pressure
vessel, i.e. RO module, are sufficient to achieve a product recovery of 40% to
50%.
An RO module with two HF elements of the type shown in Fig. 5.11 and as
described above is shown in Fig. 5.12.
The concentrate generated in the first element of the RO module flows through the
annulus formed by the outside of the elements and the wall of the pressure vessel to
the next hollow-fibre element. Here, as described above for the single element, it is
directed via the outer feed tube annulus of the double-walled tube of the second
element radially through its bundle for discharge via the module’s concentrate
collection annulus and the concentrate ports. The permeate generated in the two
elements is collected in the permeate chambers of both elements and directed via the
inner permeate tube of the module’s central double-walled tube to its product ports
for discharge.
The CTA hollow-fibre membranes are normally supplied by the manufacturer as
RO membrane modules with their elements already installed.
332 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.12 RO module with two hollow-fibre elements (Courtesy: Toyobo Hollosep)

The biggest RO membrane module for seawater desalination as supplied by


Toyobo has an external diameter of 396 mm (15.6") and length of 4720 mm
(186") with two membrane elements. This has a maximum production performance
of 95 m3/day for a salt rejection efficiency of 99.4–99.6% under the manufacturer’s
standard test conditions. When filled with liquid, it weighs around 540 kg.

5.1.3 Chemical and Mechanical Durability of Membrane Systems

The components of the RO modules of a membrane seawater desalination system,


comprising membranes, membrane elements, and pressure vessels, must, in respect
of their chemical and mechanical stability, permit reliable operation over the long
term under all operating conditions that could arise during desalination and thus
under all circumstances associated with this.
In this regard, the range of operating conditions to which the membrane system is
exposed is made up of a number of influencing factors, which are as follows:

• seawater temperature
• the compositions of the seawater and of the concentrate generated during desali-
nation with their physical and chemical properties, such as osmotic pressure, salt
content, pH, etc.
• seawater quality with its potential for fouling and scaling and the chemical
additives that are thus needed for its pretreatment before it is supplied to the
membranes, with the resulting changes in its pH and redox potential
• the nature of chemical cleaning of the membranes, the chemicals used for this,
and the cleaning conditions, such as pH, temperature, oxidative impacts, etc.
• mechanical impulses acting in particular on membrane elements due to pressure
fluctuations and build-up of differential pressure in the elements as a result of
coating of membranes.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 333

All these factors have to be taken into consideration when selecting the materials
for the membrane elements and the pressure vessels to ensure their mechanical
integrity.

5.1.3.1 Mechanical Durability


RO membrane systems for seawater desalination are operated with feed pressures at
the membrane modules ranging from 50 to 80 bar. This pressure depends on the salt
content of the seawater and of the resulting concentrate, the operating temperature,
and membrane age. The major share of the feed pressure is required for overcoming
the osmotic pressure that results as a function of the salinity of the concentrates.
Depending on the salt concentration of the feed flow to the RO unit and its product
recovery, this salinity ranges from 50,000 mg/l up to a maximum of 80,000 mg/l. For
temperatures of between 10  C and 40  C, this bandwidth corresponds to a range of
osmotic pressure from around 35–40 bar up to some 60–68 bar (see Sect. 3.2.3.3,
Fig. 3.35). A further share of the pressure is needed for overcoming the flow
resistance presented by the RO membranes with other smaller fractions to compen-
sate for the differential pressures of the elements installed in the membrane modules,
their interconnectors, and the pressure losses caused by the membrane housings.
Added to this is the differential pressure in the membrane elements that arises due to
fouling and scaling as it accumulates in the intervals between membrane cleaning
and which has to be allowed for when specifying the feed pressure.
In the second pass for subsequent permeate desalination, the feed pressures to the
membrane modules are very much lower. Depending on the permeate’s salt content
of the first pass and the operating temperature, these normally range from around
10 bar to a maximum of 15 bar. It is then the flow resistance of the RO membranes
that is the parameter that fixes the feed pressure to the membrane module.
The design of the membrane elements and their pressure vessels with the
materials selected for these must guarantee that all components of the RO unit will
reliably withstand the prevailing operating temperatures over the specified compo-
nent lifetimes.

5.1.3.1.1 Membrane Elements and Element Connectors


Operating Limits for Element Feed Pressure and Temperature
The manufacturers of TFC polyamide membranes place a maximum limit on the
pressure loading of their membrane elements for seawater desalination of 82–83 bar,
while for specific membrane types this may be only 69 bar. For the brackish water
membranes as used in the second pass of an SWRO plant, a maximum operating
pressure of up to 41 bar is permissible. Both membrane types may be operated up to
a temperature of 45  C (see Sect. 5.1.4).
However, if the elements are operated for lengthy periods in the vicinities of both
their maximum pressure and the permissible maximum temperature, this can result
in increased compaction of the separation membranes and thus degradation of their
production performance. This is then no longer reversible during subsequent opera-
tion within lower pressure and temperature ranges. For this reason, some
manufacturers stipulate that the operating pressure of their membranes be reduced
334 5 Reverse Osmosis Membrane System: Core Process of SWRO

RO SW membrane-Feed
pressure max. [bar]
85

80

75
Polyamide - TFC - Seawater membrane

70

65

60
5 10 15 20 25 30 35 40 45
Feed temperature [°C]

Fig. 5.13 Dependence of maximum RO membrane feed pressure of an RO-SW membrane on feed
temperature (Datasource: Hydranautics)

as temperature increases. Figure 5.13 shows how the permissible pressure applied to
a membrane decreases with rising feed temperature, taking as an example a
manufacturer’s TFC polyamide seawater RO membrane.
The manufacturers of CTA hollow-fibre membranes likewise quote a pressure
limitation of 82 bar and sometimes also 69 bar. Their maximum operating tempera-
ture is limited to 40  C. In this case too, the permissible maximum membrane feed
pressure decreases with rising feed temperature.
These limits on temperature and pressure stipulated by the manufacturers may not
be exceeded during operation as otherwise their performance guarantees for the
membrane elements will be voided. For this reason, an SWRO plant’s control and
monitoring systems for reverse osmosis processes must include functions that issue
alarms when these maximum values are attained and/or initiate shutting down of the
membrane systems concerned.
Additionally, the membrane elements and their interconnectors within the mem-
brane housing must be protected against the mechanical stressing that arises during
the SWRO plant’s fluctuating operating conditions, for example, start-up and shut-
down, load change, outages, etc.

Telescoping
The membrane elements and their interconnectors within the membrane housing
experience severe horizontal shock loading up to the concentrate-discharge port, in
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 335

particular when starting and during pressure build-up in the RO unit. The sudden
high-pressure fluctuations at the infeed to the RO module that act on the interlinked
spiral-wound elements and their intermediate interconnectors cause these to move
and be compressed right up to the discharge end of the pressure vessel. This is
termed telescoping. If this motion and compression are excessive, it is possible that
the connecting elements will loosen and start to leak and even the membrane
elements may suffer damage.
Telescoping may likewise arise due to fluctuations of the pressure and feed
loading of the RO modules during operation.
The end caps of the spiral-wound elements are usually so executed by the
membrane manufacturers that they reduce the telescoping action and are therefore
often also referred to as anti-telescoping devices (ATD). Additionally, to reduce or
prevent telescoping, design measures are taken at the pressure vessels as well as
precautions when they are fitted with their membrane elements.
To absorb the pressure forces exerted on the membrane housing from the
concentrate end, before inserting the membranes, a thrust ring is fitted there which
may be cylindrical as shown in Fig. 5.15, but which may also be conical in form.
After the membranes have been inserted, shimming is undertaken on the feed end
with a number of plastic shimming rings fitted at the permeate adapter. Normally, the
membrane housings are designed with a specified length tolerance so as to allow for
differences in the lengths of the membrane elements. This means that when the end
cap is fitted at the feed end, there still remains a certain play within which the
elements may move. This difference in length is packed with shimming rings so that,
as far as possible, this will no longer happen (Fig. 5.14).
Another measure to prevent pressure stressing and telescoping of the membrane
elements within the pressure vessel and thus damage to the membranes is to limit
pressure build-up with regard to both time and magnitude during start-up of the
membrane system. This means that the pressure must be ramped up over a specified
time interval and in a controlled manner. To attain such a soft start-up of an RO unit,
membrane manufacturers specify upper limits of, respectively, 20 bar/min and
30–40 bar/min for the pressure build-up rate over time for CTA HFF modules and
polyamide TFC modules. These rates have also to be maintained in the reverse case
when relieving the pressure applied to the RO modules. One TFC membrane
manufacturer limits ramping up of the feed rate to the RO system to 5%/s of the
operational feed flow.
336 5 Reverse Osmosis Membrane System: Core Process of SWRO

Shimming rings (washers)

Membrane element Pressure vessel feed end plate


assembly

Feed port

Pressure vessel feed end


permeate adapter

Fig. 5.14 Shimming on the feed side of an end-port pressure vessel

Air Venting
Before pressurizing a reverse osmosis system, any air present in it has to be vented. If
there is any air remaining in the membrane elements and the pressure vessels, they
will be placed under heavy compressive loading in the direction of flow, but also
radial stressing may arise which could substantially damage the elements and the
membrane housing. Thus, before the systems are put into operation for the first time
and whenever they are restarted, the RO components, comprising units, RO trains,
racks, etc. have to be flushed through for a sufficient length of time with pretreated
seawater at a low pressure of around 1–3 bar. In the design of the RO plant, provision
has to be made for appropriate venting pipelines and valves.

Differential Pressure of Elements and Pressure Vessel


The membrane manufacturers specify limit values for the differential pressure, which
is the feed pressure less the discharge pressure, of each element and for the total
differential pressure of the pressure vessel when it is fitted with membrane elements.
These specified values for the maximum differential pressure over each element differ
depending on manufacturer. For TFC polyamide elements, they are between 0.7 bar
and around 1.4 bar per element. The pressure loss across a pressure vessel is limited
by the manufacturers of these membranes to between 3.0 and 4.1 bar.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 337

During plant operation, pressure losses over the elements increase from their
original new and clean condition due to deposits from scaling and fouling. For this
reason, the specified limits for differential pressure must not be exceeded. If, though,
the elements are operated for too long near the maximum allowed pressure loss, due
to the then resulting substantial deposits they can usually no longer be returned to the
vicinity of the starting differential pressure by cleaning or, if so, only with great
difficulty. Also, if the limit values prescribed by the manufacturer are exceeded, the
elements could suffer mechanical damage. Thus, to ensure reliable operation, an RO
system should not be operated up to the maximum permissible differential pressure
of the membrane elements or modules, but only up to about approximately 1.15 or at
the most 1.25 times the differential pressure over the membranes when they are new
and, subsequently, after each time they have been cleaned. When they reach this
point, the membranes must be cleaned.
Regarding the differential pressure of the membrane modules, the control and
monitoring system of an RO plant should include instrumentation and monitoring
functions that register and display its current values and ensure that the prescribed
maximum values are not exceeded.

Permeate Backpressure
Mechanical damage to RO membranes may likewise occur if the pressure on their
permeate side rises above the pressure on their feed/concentrate side. In the case of TFC
membranes, delamination of the separation layer could occur, which means that it could
become detached from the supporting interlayer under the pressure on the permeate side.
Permeate backpressure may arise during shutdown of an RO unit, if, during
process standstill, the static pressure on its permeate side is higher than that on its
feed side. The reason for this could be that, due to the configuration of the piping and
the permeate buffer storage tanks, a hydrostatic pressure builds up during plant
standstill that is greater than the pressure on the RO feed side.
The manufacturers of polyamide membranes limit the static permeate-side
backpressure, i.e. the pressure difference between the permeate and the feed/concentrate
sides, to values not exceeding 0.3–0.5 bar. For CTA membrane modules, this static
differential pressure should not exceed 2.0 bar. It is possible to prevent the build-up of
such a static backpressure, i.e. negative trans-membrane pressure, through appropriate
design of the RO plant’s permeate collection and storage systems by installing either
drainage or check valves. By means of these, all the pipeline systems on the permeate
side can be drained or shut off, which, if they were to be filled, could result in a static
pressure that exceeds the limit values prescribed by the membrane manufacturers.
When the RO plant is operating, permeate backpressures that are deemed to be
dynamic are permissible, even though they exceed the prescribed static permeate
backpressure. Thus, during build-up of the pressure on the permeate side, permeate
throttling by closing valves is used when the RO plant is operating to influence the
trans-membrane pressure and so control permeate output. If the process is controlled
in this way, however, it must be ensured that the material on the permeate-side
interconnectors and the permeate ports of the membrane housing can withstand the
pressure loading for the given RO operating temperatures.
338 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.1 Temperature Maximum dynamic


dependence of maximum Temperature ( C) backpressure (bar)
allowable dynamic
20 23.3
backpressure of pressure
vessels with spiral-wound 25 20.6
elements 30 17.7
35 15.1
40 12.4
45 10.0
For CTA membrane modules, the dynamic permeate backpressure is
limited to 5 bar

Values for the maximum permissible dynamic permeate backpressure of mem-


brane pressure vessels with spiral-wound membrane elements depending on feed
temperature are compiled in Table 5.1.5

Permeate Drawback
During outages of a seawater desalination RO system, due to the large difference in
salt concentration between the two membrane sides, when the pressure is removed,
forward osmosis occurs (Fig. 5.1), i.e. permeate water flows from the permeate side
of the membranes back to the feed/concentrate side. This “drawback” or “suck-back”
can result in dehydration of the membranes if there is not a sufficient volume of
permeate available for its backflow through the membranes. This could have the
consequence of reducing membrane performance. Under unfavourable conditions, a
vacuum may form and also air may collect on the pressure side of the membrane
elements. Undisturbed permeate backflow, though, could certainly support the
membrane cleaning action.
If an SWRO plant is shut down, its seawater membrane system is flushed with
seawater or, even better, permeate, and due to the reduction of the salt content on the
feed and concentrate sides, permeate backflow may be reduced or completely
prevented. This flushing operation is an important first step when taking an RO
seawater desalination plant out of service, not only to reduce or avoid drawback but
also to protect against shutdown corrosion of the metal components on the seawater
and concentrate sides as well as the formation of deposits on the membranes. For
reliable operation of the RO plant, it must therefore be ensured that this flushing
process be carried out for scheduled as well as unscheduled outages. This is thus a
vital step in the process of taking a seawater desalination RO plant out of service, so
it should be automated and hooked up to the plant emergency power supply.
An additional safeguard to cover the case of an unscheduled outage and to
prevent damage to the membranes due to inadequate drawback is to install a
drawback or suck-back tank in the RO permeate collection system. However, this
tank will only be needed if the volume remaining in the permeate pipelines during

5
Datasource: DuPont Filmtec Reverse Osmosis Membranes Technical Manual Table 30.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 339

plant outage is not sufficient to permit adequate drawback. The volume of the
drawback tank is calculated as follows (Eq. 5.1):

ðV ME  nES Þ  V Pp
V DBT ¼ ð5:1Þ
1000
VDBT ¼ volume of drawback tank [m3]
VME ¼ volume of water in membrane element [l] ¼ 25 l for 8" element
VPp ¼ volume of permeate piping between pressure vessels of RO and drawback
tank [l]
nES ¼ number of membrane elements installed in the RO unit []
The drawback tank must always be filled and shall be so arranged that, during an
outage, the maximum permissible permeate backpressure will not be exceeded. This
means that, for polyamide membrane elements, it should not be mounted more than
3 m above the bottommost membrane module tier of an RO rack.

5.1.3.1.2 Membrane Housings


Dimensions and Maximum Operation Pressure
Determining factors for dimensioning the pressure vessel for an RO unit, i.e. its
internal diameter and length, are the size and number of membrane elements that are
installed within it. The maximum operating pressure that the housing has to with-
stand results from the operational conditions of the RO unit and its membrane
elements as well as the influences that have to be taken into consideration when
configuring the membrane modules within the system, comprising safety
allowances, fouling, scaling and ageing of the membranes as well as safety factors
that are specified to take into account abnormal plant conditions.
Because the polyamide TFC spiral-wound elements of virtually all membrane
manufacturers are manufactured in standardized dimensions with regard to element
length and gradation of diameters, for this membrane type the membrane housing
can be standardized to a certain extent. This likewise applies for the pressure ratings
of the housing as the feed pressures specified for the respective membrane element
types by the membrane manufacturers are all of approximately of the same order.
In line with the standardized diameters of the 4", 8", and 16" membrane elements
of ~101, ~201, and ~402 mm, respectively, pressure housings for polyamide
membrane elements also have gradations of their internal diameters to match the
diameters of these elements. The TFC spiral-wound elements used in industrial and
communal desalination plants are approximately 40"–41" (1016–1039 mm) long. In
these plants, depending on the capacity of the RO system and its configuration, the
membrane housings are loaded with up to 6–8 elements. Accordingly, the pressure
vessels have lengths of 6–8 times the length of the membrane elements plus
additions as appropriate to accommodate the membrane element interconnectors
and end caps. For lower-rated plants, also available are 2.5" and 4" elements with
standardized shorter lengths.
340 5 Reverse Osmosis Membrane System: Core Process of SWRO

This standardization of the dimensions of polyamide TFC elements has of course


the advantage that existing pressure vessels can also be fitted with membrane
elements from different manufacturers.
Manufacturers offer their membrane pressure vessels in a number of gradations of
maximum permissible operating pressure. In the high-pressure range, i.e. for RO
seawater desalination systems, the available gradations are 1500 psig (~103 bar),
1200 psig (~83 bar), and 1000 psig (~69 bar), and in the intermediate and
low-pressure ranges, these are 600 psig (~41 bar) down to 300 psig (~21 bar). For
seawater desalination systems, primarily pressure vessels are used that are rated at
1000 psig and 1200 psig, whereas for permeate post-desalination, i.e. the second
pass, 600 psig vessels find application. These maximum permissible operating
pressures apply for temperatures of up to 49  C, while some manufacturers also
quote temperatures of up to 66–69  C.

Quality Assurance in Housing Construction and Fabrication


For seawater desalination, but also for industrial and communal brackish water and
fresh water desalination, almost exclusively membrane pressure vessels are used
whose housings are made of fibre reinforced plastic (FRP), mostly glass-fibre
reinforced epoxy resin. The remaining pressure vessel parts that are in contact
with seawater or concentrate, like seals, permeate tubes, permeate adapters, permeate
ports, thrust rings, and sealing plates of end caps, are primarily of plastic materials.
Only the retaining rings or, more precisely, their segments with which the end caps
are fixed in the retaining ring grooves of the pressure vessel together with their
attachment elements and the base or backing plates are of metal.
For pressure vessels for seawater desalination and vessels at intermediate pressure
as used in the second pass of an SWRO plant, the ports at the housing feed and
concentrate ends are of metal.
These pressure vessels are offered in three designs, namely, end-port, side-port,
and multi-port versions, to match their feed and concentrate discharge configurations
(also see Sects. 5.5.2 and 5.5.2.1).
An example of the design of an end port pressure vessel with matching
components that make up the end caps and the vessel internals is shown in
Fig. 5.15. Depending on pressure vessel manufacturer, these parts may differ in
their design details, but for all the vessel designs principle is the same.
The high pressures prevailing in a seawater desalination system place high
demands on the quality of design and manufacture of the membrane housings.
Weak points in the structure of the housing walls and the design and sealing of the
end caps could considerably endanger operating crews and may result in accidents
with costly material damage. Leaks at the end caps of the housing with egress of
seawater or concentrate could cause severe corrosion of plant components. For this
reason, the basic condition for safe operation over the long term of a high-pressure
membrane system is that the membrane pressure vessel be designed and
manufactured to an appropriate high quality. Likewise of importance for the
membrane’s long-term pressure resistance is that, with regard to their design and
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 341

101 121
111
81 91
71

61
51
41

31
21

11

No. Part No Part


1 Retaining ring 7 Concentrate port
2 Backing plate 8 Endplate seal
3 Sealing plate 9 Permeate tube
4 Thrust ring 10 Retaining ring groove
5 Brine seal of element 11 Permeate adapter
6 Membrane element 12 Pressure vessel body

Fig. 5.15 End cap construction of end-port pressure vessel

choice of materials, appropriate safety factors are specified to take account of


material ageing.
Pressure vessels are available in coded or non-coded versions. For the coded
vessels, the design, selection of materials, manufacturing process, and qualification
tests of prototypes of vessel groups or series and hydrostatic leakage tests are based
on recognized standards and guidelines that are largely independent of
manufacturers. Such sets of standards and regulations are:

• The Boiler and Pressure Vessel Code (BPVC), Section X: Fiber–Reinforced


Vessel Plastic Pressure Vessel der American Society of Mechanical Engineers
(ASME)6

6
ASME Boiler and Pressure Code (BPVC) Section X-Fiber-Reinforced Plastic Pressure Vessels—
Latest issue.
342 5 Reverse Osmosis Membrane System: Core Process of SWRO

• Directive 2014/68/EU7—The European Pressure Equipment Directive (PED)


with the implementing regulations of the EU member states as well as the
standards that have been harmonized within the EU.

Both sets of standards and regulations, i.e. the ASME BPVC Code and the
European Pressure Equipment Directive, together with the ordinances of the EU
member states and the harmonized EU standards, contain guidelines and
specifications for the design, manufacture, and quality control of pressure vessels.
In both the USA and Canada, the ASME BPVC Code is the generally accepted and
applied set of standards. But also internationally, it finds application in numerous
countries as the quality standard for industrial, communal, and government projects
and institutions. This likewise applies for Section X of the ASME BPVC code for the
manufacture of pressure vessels of fibre reinforced plastic (FRP) and thus also for the
production of RO membrane pressure vessels of this material. The ASME BPVC
Code X has become established worldwide for membrane desalination technology as
the quality basis for the manufacture of such pressure vessels. This is a comprehen-
sive set of standards and contains obligatory technical rules and specifications for
manufacturing FRP pressure vessels with regard to their dimensions and design, the
type of materials used, and how manufacture is to proceed as well as for the
procedures for quality control of the products and the testing methods to be
employed for this purpose.
So as to be able to document that its products have been manufactured in full
conformity with the ASME Code X, a manufacturer has to have a number of quality
assessments performed by independent certification institutions that have been
accredited by the ASME, referred to as Authorized Inspection Agencies (AIAs), at
the various manufacturing stages. These include:

• Manufacturer accreditation: To obtain this, for the certification procedure the


manufacturer must show that he is capable of complying with the ASME X Code
stipulations throughout the entire manufacturing process and can also monitor
this by means of a quality management system. To this end, he must compile a
quality assurance manual that takes into account all requirements laid down in the
Code and must implement this QA system.
• Design qualification: Evidence shall be presented to a certified professional
engineer, i.e. an authorized inspector (AI) of the AIA, that the design,
manufacturing process, and materials comply with the specifications of the
ASME Code.
• Quality verification of the vessels witnessed by an AI: These include:
– qualification tests: In order to obtain a qualification certificate for vessel
groups or series, among others, prototype vessel cycling tests are conducted

7
Directive 2014/68/EU of the European Parliament and of the Council of 15 May 2014 on the
harmonisation of the laws of the Member States relating to the making available on the market of
pressure equipment (European Pressure Equipment Directive PED).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 343

to demonstrate that they can withstand a qualification pressure that normally


corresponds to six times the design pressure over a specified number of
pressurization-depressurization cycles.
– hydrostatic leakage tests of the completed pressure vessels at 1.3 times the
design pressure without the occurrence of leaks.

If the pressure vessels are in compliance with the requirements of the ASME
Code X with regard to design, manufacture, and pressure tests, this is documented
by an authorization certificate and the finished products may be labelled accord-
ingly. In the case of the ASME Code, this is a special stamp. If it is required that
during the production process specific manufactured products be labelled with
this stamp, then an additional AIA inspection during component production is
needed.
The ASME certificate of authorization is issued specifically for the inspected
manufacturing facilities of a company. It is issued for a specific time period after
which it has to be renewed.
The European Pressure Equipment Directive (PED) 2014/68/EU covers pres-
sure equipment and assemblies with a maximum permissible pressure that exceeds
0.5 bar that are to be put onto the market in EU member countries. It applies, among
others, for metallic and non-metallic pressure vessels as well as pipelines. For the
design and manufacture as well as the materials of these products, the PED lays
down specifications for fulfilling certain essential safety requirements (ESR) (PED
Annex I, Essential Safety Requirements) that are mandated for such pressure
equipment. Requirements are specified for:

• design and manufacturing quality as well as their quality monitoring


• equipping the products with safeguards
• product identification
• documentation of the manufacturing processes
• provision of instructions for their installation and operation.

In order to check, determine, and document the conformity of the product with the
PED specifications, the directive stipulates that a conformity assessment procedure
be followed. The scope of this procedure depends on the medium and the pressure
with which the pressure vessel is operated and what its volumetric capacity is. In line
with these criteria and the product of pressure and volumetric capacity of the vessel,
it is assigned to a PED hazard category. From this and with the help of an assignment
graph (PED Annex II Conformity Assessment Tables), there results a module
classification and, depending on the type of module, the scope and the testing
procedures (PED Annex III Conformity Assessment Procedures) for the vessel’s
design, manufacture, and quality control are set forth. The higher the hazard cate-
gory, the greater the outlay for inspection and documentation for the conformity
assessment procedure. Whether an independent certification institute witnesses or
conducts the test as “notified body” likewise depends on the hazard category and the
module that results from this. The lowest hazard category is the range specified in
344 5 Reverse Osmosis Membrane System: Core Process of SWRO

Article 4 Paragraph 3 of the PED, for which sound engineering practice (SEP) is
stipulated. For this, no design and production inspection procedure are needed.
In Guideline No. 9.1, the EU8 PED guidelines define “sound engineering prac-
tice” as follows:

“Sound engineering practice means that such pressure equipment is designed taking
into account all relevant factors influencing its safety. Furthermore, such equip-
ment is manufactured, verified, and delivered with instructions for use in order to
ensure its safety during its intended life, when used in foreseeable or reasonably
foreseeable conditions. The manufacturer is responsible for the application of
sound engineering practice.
SEP is generally taken to mean the use of industry standards and practices that have
been in use in the member states of the EU for many years and have demonstrated
their safety over time. This does not exclude the use of novel or innovative design
and manufacturing processes, provided that the processes are based on sound
engineering principles, including adequate testing of all materials and
processes.”

In the next higher category, Module A, too, the manufacturer need only undertake
a company-internal inspection and issue a declaration of conformity. For the
subsequent higher modules A1, B, C1, D, and D1, the check of conformity is
undertaken by a “notified body” company. For all categories, following completion
of the inspections and on the basis of the issued EC Examination Certificate, the
manufacturer issues a declaration of conformity for his product. Thereupon, his
product is issued with the CE mark to document its conformity with the PED.
Products that are assigned to the SEP range are not issued with a CE mark.
For RO pressure vessels with internal diameters of 4" (~101 mm) to 16"
(~402 mm) and housing lengths of up to ~8000 mm that fix the volumetric capacity
as well as operating pressures of between 41 and 83 bar, there result classifications
within the SEP range and in Hazard Class 1 corresponding to Module A of the
classification scale.
However, the manufacturer of a pressure vessel can select a higher class for the
configuration examination if it is important for him that the design, manufacture, and
quality control of his product be evaluated under a more rigorous examination
procedure and that this then also be documented in the EC Examination Certificate.
However, the CE mark with which he is issued will only be in accordance with the
hazard class to which his product is assigned.
As the PED is in the nature of a framework directive, in the main it consists of
general and functional stipulations and, to a substantially lesser degree than is the
case for the ASME Code X, detailed technical rules for conceptual and detail design
as well as manufacture. A set of technical standards that is comparable with the
ASME Code results from the PED together with the standards issued for the

8
http://ec.europa.eu/enterprise/sectors/pressure-and-gas/documents/ped/guidelines/index_en.htm
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 345

respective pressure vessel that have been harmonized in the EU member countries.
Finding application for pressure vessels of glass-fibre reinforced polyester resin are
the standards EN 13923 Filament-wound FRP pressure vessels—Materials, design,
manufacturing and testing9; and EN 13121 Parts 1–4 GRP tanks and vessels for use
above ground.10
Although the design, material selection, and manufacture of a pressure vessel on
the basis of harmonized EU standards facilitate proof of its conformity with the PED,
the application of EN standards is not compulsory. Thus, a membrane pressure
vessel designed and manufactured on the basis of the ASME Code X or other
standards that are in accord with sound engineering practice may be issued with a
certificate of conformity following a check of such conformity together with any
adaptations to the PED requirements that may be necessary. However, in this case
the basis for the technical design and the materials, the manufacturing process, and
product quality assurance has to be presented and documented in detail in the course
of the check of conformity.
So-called “non-coded” pressure vessels are produced on the basis of the
manufacturer’s own design and manufacturing standards. It is certainly possible
that these internal standards will take as a starting point the aforementioned ASME
and/or EN standards. However, there will be no independent third-party inspection
of the design, materials, manufacturing process, and quality control of the products.
If these products are to be sold on the EU market, they will require a check for
conformity with the PED, any adaptations that are needed to conform with the PED,
and issue of a declaration of conformity by the manufacturer, as described in the
above.

5.1.3.2 Chemical Durability


A primary requirement for all membrane systems used for seawater desalination is
that the material of the membranes themselves as well as the materials of the
membrane elements and the pressure vessels in which the elements are mounted
are resistant to corrosive attack by seawater and by the concentrates that are
produced during desalination. A major part of the materials and plastics used in
RO modules are plastics that meet these stipulations. However, the modules’
pressure vessels contain metallic components, such as sealing rings, end plates,
and bolted connections as well as feed and concentrate-discharge ports that are in
contact with seawater or concentrate at times or continuously. These, therefore, have
to be manufactured of stainless steel to an appropriate quality, such as Super Duplex
or highly alloyed austenitic steel (see [13] Chap. 6 in Volume 2). But, additionally,
these materials have to be resistant to the conditions during chemical cleaning of the
membranes and also oxidizing influences that may arise during the desalination
process.

9
EN 13923 Filament-wound FRP pressure vessels—Materials, design, manufacturing and testing.
10
EN 13121–1-4 GRP tanks and vessels for use above ground—Part 1: Raw materials; Specifica-
tion conditions and acceptance conditions; Part 2: Composite materials—Chemical resistance; Part
3: Design and workmanship; Part 4: Delivery, installation and maintenance.
346 5 Reverse Osmosis Membrane System: Core Process of SWRO

A determining factor for the operating parameters of an RO system is in particular


the material of the separation layer of its membranes, i.e. either CTA or polyamide.
Their chemical resistance sets limits for operation of membrane desalination and
compliance with these limits is critical for the long-term behaviour of the membrane
elements.
The two types of membrane used for seawater desalination are as follows:

• aromatic polyamide in the form of TFC membranes


• cellulose triacetate (CTA) as hollow-fibre membranes

They exhibit significant differences in their resistance to certain pH ranges and their
sensitivity to oxidation, especially when exposed to chlorine. This means that the
range of chemical operating conditions differs between RO modules with CTA HF
membranes and those with TFC polyamide membranes.
Oxidative influences on the separation layer of TFC polyamide membranes and of
CTA membranes can arise with the application of inorganic disinfectants as used for
preventing biological growth and biofouling during seawater pretreatment and on
the membranes of RO systems. Such disinfectants are in particular chlorine and
chlorinated compounds in the form of:

• gaseous chlorine, Cl2


• sodium hypochlorite solution, NaOCl
• chlorine dioxide, ClO2
• chloramine, NH2Cl.

In desalination plants fed with surface water, chlorine is the most widely used
disinfectant and biocide to combat biological growth and biofouling in their RO
systems. Because of its severe membrane damaging properties, it has to be largely
removed again before the treated water is fed into RO systems, in particular if they
are equipped with polyamide membranes. Efforts are therefore being made to find
alternative disinfectants for such plants. These are the above-named compounds,
chlorine dioxide and chloramine. The changes to the chemical and physical structure
of the separation membrane layer due to the action of the various chlorinated
disinfectants take place under a range of chemical mechanisms.
When drinking water and process water from fresh water sources are supplied to
conventional water treatment processes, ozone is widely used as disinfectant. Com-
pared with chlorine, though, this has a still higher oxidation potential. Thus, RO
membranes may suffer severe and irreversible damage after even just a relatively
short contact with this disinfectant. This is a major reason for ozone finding hardly
any application up to now as a biocide in membrane desalination technology. A
further factor for seawater desalination is that ozone oxidizes the bromide present in
seawater to bromate to a greater extent.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 347

5.1.3.2.1 Chemical Impacts on Polyamide Membranes


Chlorine: Halogenation and Oxidation Impacts and Chlorine Tolerance Values
Seawater contains between 50 and 80 mg/l of bromide subject to its salt content.
Depending on the form in which chlorine is dosed, i.e. as Cl2 or OCl, and the pH of
the seawater to which it is fed, it reacts with the bromide ions of the seawater as
shown in the following chemical equations to form bromium, hypobromous acid, or
hypobromite:

Cl2 þ 2Br ⇄Br2 þ 2Cl

Cl2 þ H2 O⇄HOCl þ Hþ þ Cl

HOCl þ Br ⇄HOBr þ Cl

HOCl⇄OCl þ Hþ

OCl þ Br ⇄OBr  þ Cl

At a low pH in the acidic range, chlorine hydrolyses to an increased extent to


hypochlorous acid which then, with rising pH, dissociates to hypochlorite. For the pH
feed range usual for seawater desalination from a minimum of 6.0 (with acid dosing
to combat carbonate scaling) to a maximum of 8.0 to 8.5, dosed chlorine or
hypochlorite is therefore present as a mixture of hypochlorite and hypochlorous
acid that then reacts in this form with the bromide present in the seawater.
The bromine formed by the addition of chlorine to the seawater hydrolyses, in the
acidic pH range, to hypobromous acid HOBr, which then in turn forms in part or
completely hypobromite OBr.

Br2 þ H2 O⇄HOBr þ Hþ þ Br

HOBr⇄OBr þ Hþ

Whereas up to a pH of 6 almost exclusively hypobromous acid HOBr is present,


as pH rises from the neutral value of 7, HOBr increasingly dissociates to
hypobromite OBr.
The chemistry of seawater chlorination is described in detail under [13] Chap. 2,
Sect. 2.2.1 in Volume 2, which also shows the distribution curves of chlorine and
bromine compounds as functions of pH.
Regarding the proportion of bromine that is still present, investigations
undertaken at thermal seawater desalination plants have shown that, even with
seawater chlorinated into the alkaline range with a pH of 8–9, up to 5% of the
bromide content present in seawater is still detectable as free bromine, and this share
rises with increasing temperature [14].
For chlorine dosing rates of 2–10 mg Cl2/l (depending on whether dosing is
continuous or intermittent), for an average bromide content of the seawater of 65 mg/
l Br, this corresponds to a maximum molar ratio of chlorine to bromide of ~0.17
348 5 Reverse Osmosis Membrane System: Core Process of SWRO

mole Cl2/mole Br. This means there is a high excess of bromide present and that the
dosed chlorine is almost completely transformed in the oxidation reactions with it
[15]. Under kinetic aspects, too, the HOCl/OCl/Br reactions proceed very rapidly,
taking only a few seconds [16], so that of the dosed chlorine there is hardly any still
available after just a short time. This also applies of course for the oxidative
influence on the material of the membrane separation layers. Although this dosing
of chlorine to the intake pipeline of an SWRO plant is thus the trigger for potential
oxidative influencing of the structure of the RO separation membranes, the chemical
compounds that actually directly impact the structure of the membrane separation
layer are the oxidizing and halogenating bromine compounds that are formed by the
injection of chlorine.
It may be assumed that upon contact of the membrane separation layer with
chlorine or the bromine generated from bromide by secondary reactions, the impact
on the chemical structure of the aromatic polyamide due to the active halogen of the
disinfectant is such that, in a first reaction, the nitrogen of the aromatic rings of the
amide groups connecting the polyamide to N-chloroderivatives is halogenated.
There then follows halogenation of the aromatic rings of the polymer itself with
the halogen attaching itself to these or having a substituting action (Fig. 5.3) [17]. In
parallel with this, further oxidizing and hydrolyzing reactions proceed which could
likewise modify the polymer’s chemical structure. This could result in cleavage of
the amide bonds and also to breakdown of those bonds that serve as cross-linking
elements between the polyamide chains.
Which of these reactions contributes most to modifying the molecular structure of
the polyamide depends on the one hand on its chemical structure but, on the other, on
the pH prevailing during the action of the halogen-containing biocide on the
membranes. The latter is logical as, depending on the pH and whether the halogen
is Cl2 or Br2, halogens are present in differing compound forms. For both halogens,
the related hypohalogenous acid—HOCl or HOBr—in particular is an especially
effective halogenation agent. Because for bromium in the acidic or weakly alkaline
pH range, active halogen is present mainly or to a large part in the form of
hypobromous acid HOBr; it may be inferred that in this range halogenation is the
predominant reaction. But elemental bromium has also to be considered, as it still
arises to a noticeable extent in this pH range, and it may act as both a halogenation
and an oxidizing agent.
Oxidizing and hydrolyzing processes occur primarily in the strongly alkaline
range by interaction of the existing alkalinity with the halogen compounds there
present in the form of hypochlorite OCl or hypobromite OBr [18].
All these reactions speed up if the temperature is raised. Through the catalytic
action of metals, like iron, manganese, zinc, copper, aluminium, etc., the impact of
halogens on polyamide membrane separation layers is much intensified.
Apart from chemical changes in the molecular structure of the polyamide, the
membrane separation layer is also greatly affected in its physical and mechanical
properties. Thus, the surface charge of the membranes and the degree of their
hydrophilicity or hydrophobicity may change. At the same time, physical surface
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 349

structures are also modified, for example, roughness and porosity as well as, with
increasing chemical action, their mechanical stability in general.
The consequence of these chemical and mechanical changes in the membrane
structure is that the salt rejection capability of the membranes decreases and their
product flux properties change. As the action of halogens increases, above all their
salt rejection capability drops. Particularly pronounced is the reduction of salt
rejection efficiency for monovalent ions. Depending on membrane type and
operating conditions in terms of pH, i.e. in the acidic, neutral, or weakly alkaline
range, the membranes’ permeability may increase or decrease.
The influence of chlorine on RO membranes has been tested in numerous
investigations, primarily in the laboratory, and the results have been published. In
these, the degree to which a membrane is influenced by the variable CTmembrane,
i.e. by the product of the concentration in contact with the membrane cc,halogen and
the contact time at this concentration, defines τdisinfectant (Eq. 5.2).

CTmembrane ¼ cc,halogen  τc,halogen : ð5:2Þ

CTmembrane ¼ concentration/contact time product of membrane [mg/lh; ppmh]


cc,halogen ¼ concentration of halogen in contact with membrane [mg/l, ppm]
τc,halogen ¼ contact time of halogen in contact with membrane [h]
For these test series, values for CTmembrane were selected which made it possible,
by dosing chlorine at an increased rate, to determine its action on the membranes, as
far as this could be done over a defined, limited testing period. However, these
dosing rates do not normally correspond to the substantially lower dosing
concentrations used in practice for disinfecting seawater in the infeed to SWRO
plants.
For the action of bromine on desalination membranes which predominates for
seawater desalination, only limited trials results are available. Most of these experi-
mental trials were conducted similar to those for chlorine [19–21]. Investigations
under conditions encountered in practice into the action of bromine on seawater
desalination membranes should, though, be conducted with the same molar ratio of
injected chlorine to the bromide present in seawater of a maximum of 0.2 mole/mole,
to match the operation of SWRO plants.
Trials to determine the action of bromine on RO membranes have shown that this
is markedly different from that of chlorine. This is attributable in part to the differing
dependency on pH of bromine hydrolysis in comparison to that of chlorine and the
dissociation of the hypobromous acid, but also to the differing oxidation potentials
of the various bromine components compared with chlorine compounds. Regarding
the degrees of halogenation and oxidizing action on an FT-30 polyamide membrane,
from the investigation results in [22] for chlorine and bromine and their compounds,
the following ranking can be derived for these membranes:

OBr > OCl > HOBr $ Br2 > Cl2 $ HOCl


350 5 Reverse Osmosis Membrane System: Core Process of SWRO

In the alkaline pH range in particular, membrane performance is greatly changed


due to hypobromite OBr. But also with neutral and acidic pH, the action of
hypobromous acid and bromium is greater than is the case with the chlorine
compounds. For other polyamide membrane types, in contrast there is an increased
impact of the halogen on the membrane separation layer more in the acidic range due
to the predominance of hypochlorous acid HOCl in this range or, in seawater,
primarily bromine Br2 and hypobromous acid HOBr.
The chemical and physical changes at the polyamide separation layer are irre-
versible and accumulate over the time for which they are in contact with the
halogenating and oxidizing medium. The influence of the medium is not solely
determined by the concentration of the disinfectant it contains, but also by the
contact time of the membranes with the compound, and ultimately by the product
of these, CTmembrane (Eq. 5.2).
The short-term impact of elevated halogen concentrations on a membrane need
not, therefore, lead to a noticeable impairment of their salt rejection rate nor a change
in their production performance. If, though, such operating conditions arise more
frequently, the impacts of all such events on the membranes accumulate accordingly
and result in a corresponding reduction of their potential operational lifetime.
As a tolerance range for the action of chlorine on polyamide TFC membranes,
membrane manufacturers quote CTmembrane values of 200 to 1000 ppm Cl2h up to
doubling of the salt transport through the membrane. For a TFC seawater desalina-
tion membrane, this means that then the original salt rejection rate of the membranes
of 99.8% has been reduced to 99.6%.
This tolerance range, however, should only be regarded as a rough guide, but it
clearly shows the wide spread of these values. The CTmembrane value, though, allows
a quantitative comparison to be made between different types of separation
membranes in terms of their response to chemical action and their sensitivity to this.
As described above, the degree of impact of halogens and oxidants on polyamide
membranes may substantially increase at elevated temperatures and pH as well as
due to the catalytic action of heavy metals and thus also drastically reduce the
tolerance limit. Also, of course, the type of polyamide material used for the separa-
tion membranes influences their halogen tolerance.
Due to the increased halogenating and oxidizing action of bromine, hypobromite,
and hypobromous acid on polyamide membranes during seawater desalination, the
halogen tolerance limit should be assigned more to the lower section of the above
named tolerance range, that is to below 500 ppm Cl2h. However, at present there are
no measurement data available for the conditions of membrane desalination plants
encountered in practice that would allow a direct comparison of the influence of
bromide content on polyamide membranes for chlorination under seawater
conditions with the situation for chlorinated freshwater with just a low bromide
concentration.
To enable an estimate to be made, with the help of such tolerance values, of the
influence of halogen on the service life of a polyamide membrane under the
operating conditions of a seawater desalination plant encountered in practice, the
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 351

relationship shown under Eq. (5.2) would have to be adapted to match the actual
operating conditions of such a plant.
Over a lengthy operating period, the dosing concentration of a halogen or halide
used for disinfection exhibits fluctuations. The calculation is therefore to be based on
the average contact concentration cØ,c,halogen (Eq. 5.3):

CTmembrane ¼ c∅,c,halogen  τc,halogen : ð5:3Þ

cØ,c,halogen ¼ average concentration of halogen in contact with membrane [mg/l,


ppm]
This average concentration is calculated from a corresponding number of indi-
vidual values of the halogen concentration ci,c,halogen at the specific dosing conditions
(Eq. 5.4):

nP
i¼x
ci,c,halogen
ni¼1
c∅,c,halogen ¼ : ð5:4Þ
ni¼x
ci,c,halogen ¼ concentration of halogen in contact with membrane at dosing case ni
[mg/l, ppm]
ni ¼ number of dosing cases []
It must also be taken into account that the dosing concentration ci,d,halogen is
reduced on the path from the dosing location to the membranes by Δi,halogen owing to
the consumption of halogen by water constituents or because dechlorination takes
place upstream of the membranes. This means that the membrane is only exposed to
the residual content ci,c,halogen. The extent of this chlorine, or also halogen, attrition
Δi,halogen, depends on the water conditions and operating situation during dosing,
i.e. presence of reducing constituents in the water, temperature, pH, etc.

ci,c,halogen ¼ ci,d,halogen  Δi,halogen : ð5:5Þ

ci,d,halogen ¼ dosing concentration of halogen at dosing case ni [mg/l, ppm]


Δi,halogen ¼ halogen consumption by water or dechlorination at dosing case ni
[mg/l, ppm]
Taking the mean halogen contact concentration that is so obtained, the maximum
allowable contact time τc,halogen max for the polyamide membrane concerned is
calculated using Eq. (5.6). This time value, though, corresponds to the anticipated
service life of the membrane only if it is continuously exposed to this halogen
concentration.

CTmembrane
τc,halogen ¼ τc,halogen ¼ : ð5:6Þ
max
c∅,c,halogen
352 5 Reverse Osmosis Membrane System: Core Process of SWRO

τc,halogen ¼ contact time of halogen in contact with membrane [h]


τc,halogen max ¼ maximum allowable membrane contact time with halogen of cØ,c,
halogen [h]

In practice, for disinfection of membrane desalination plants, the disinfectant is


usually dosed intermittently. The ratio of the total duration of dosing to the plant
operating time under consideration must then be accounted for by the factor fct
(Eqs. 5.7 and 5.8).

τc,halogen max
τMOp ¼ : ð5:7Þ
max
f ct
nP
i¼x
τi, dosing
ni¼1
f ct ¼ : ð5:8Þ
τO
τMOp max ¼ maximum membrane lifetime in contact with halogen of cØ,c,halogen
[h]
fct ¼ halogen dosing time to operation time factor [h/h]
τi,dosing ¼ duration of dosing case ni [h]
τO ¼ duration of operation period with nx dosing cases [h]
Because in this case the total duration of the individual dosing cases is always less
than the plant’s operating time τO, i.e. the quotient fct of these is <1, there then results
a maximum membrane service life τMOp max, which is greater than the allowable
membrane contact time τc,halogen max.
If to obtain guide values for the polyamide membrane halogen contact time for
seawater desalination, a value of 500 ppm Cl2h is taken and assuming a chlorine
dosing concentration of 5–10 mg Cl2/l for intermittent dosing for dechlorination is
reduced to a mean value of 0.05 mg/l then, in accordance with Eq. (5.6), the
maximum allowable contact time τc,halogen max is calculated as 10,000 h, equivalent
to 1.14 years, provided the membranes are continuously exposed to this mean
chlorine concentration. With values for fct of 0.04 to 0.1, corresponding to intermit-
tent chlorine dosing of around 2–5 h per 48 h, applying Eq. (5.7) results in a
maximum membrane service life referred to halogen contact of 11.4–27.3 years.
This shows that for polyamide membranes, intermittent chlorine dosing and
corresponding dechlorination are necessary to attain membrane service lives of
more than 5 years, which is specified as the state-of-the-art target for membrane
desalination plants. However, the service life of RO membranes is not solely
determined by the action of inorganic disinfectants, but is also influenced by
numerous other factors (see Sect. 5.1.4).

Impacts of Chlorine Dioxide and Tolerance Values


Chlorine dioxide is a gas with strongly oxidative and biocidal properties comparable
to those of chlorine and it is readily soluble in water. ClO2 does not hydrolyse in
water, and thus, at least in the acidic, neutral, and weakly alkaline range, its oxidizing
action, unlike that of chlorine and hypochlorite, is influenced hardly at all by pH.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 353

In the alkaline range with a pH > 9, this compound disproportionates to chlorite


and chlorate.

2ClO2 þ 2OH ! ClO2  þ ClO3  þ H2 O

In the acidic range, disproportioning can likewise occur with the additional
formation of chloride.

4ClO2 þ 2Hþ ! 2ClO2  þ ClO3  þ Cl þ H2 O

These disproportioning reactions are catalysed and accelerated by exposure to


light, at elevated temperature and in the presence of heavy metals.
Chlorine dioxide solutions are not stable over longer periods and they also cannot
be transported, even if catalysing actions are avoided, so the dosing solutions must
be prepared directly at their place of use.
When chlorine dioxide contaminated with chlorine—as is generated during the
chlorite/chlorine process that finds widespread application for water treatment—
comes into contact with a polyamide TFC membrane, this is oxidized and
halogenated, leading to rapid damage of its structure and a substantial deterioration
of its salt separation and transmission properties. This mixture of oxidizing agents
then attacks the membrane even more than would be the case if it were to come into
contact solely with chlorine.
Unlike the action of chlorine and bromine, pure chlorine dioxide has only an
oxidizing action and does not halogenate polyamide TFC membranes, so the damage
to the membrane separation layer is substantially less than that caused by the
halogens. This was revealed by a number of membrane tests with ClO2 that had
been produced so as to contain no chlorine.
The Fujiwara test and additional structural analyses like x-ray spectroscopy
undertaken at separation membranes treated over a lengthy period with pure chlorine
dioxide solutions not contaminated with chlorine showed that there were no deposits
of halogens to be found in the polyamide structure of the membrane separation
layer [19].
During other tests, brackish water and seawater polyamide membranes were
treated over a lengthy period on a lab scale with chlorine dioxide solutions at
concentrations of 1–10 mg ClO2/l [22, 23] and up to over 100 mg ClO2/l [24, 25]
while under operating conditions in membrane desalination plants, membranes were
exposed to concentrations in the range of 0.5–2 mg ClO2/l [26]. It was found that,
under certain conditions, chlorine dioxide causes substantially less deterioration of
the properties of polyamide TFC membranes of type FT-30 than is the case with
chlorine. Additionally, operational trials showed that this compound has a
favourable biocidal and disinfecting action that is comparable to chlorine.
Under the conditions of the contact tests on lab scale [22, 24, 25], CTmembrane
values for chlorine dioxide of around 10,000–30,000 mg ClO2/lh were determined.
However, in line with the findings of the tests and operational trials conducted up to
now, such values for the compatibility of polyamide TFC membranes with chlorine
354 5 Reverse Osmosis Membrane System: Core Process of SWRO

dioxide can only be attained if the contact between ClO2 and membrane is in the
weakly acidic to the weakly alkaline range, i.e. for a pH from around 6.0–8.0. At a
higher pH, the membrane suffers more rapid damage. Heavy metals, like iron,
manganese, and copper, catalyse the oxidizing action of chlorine dioxide, which
means they reduce the CTmembrane value of a membrane.
Normally, the required dosing rate for disinfection with ClO2 is lower than if
chlorine is used. To prevent biological fouling when extracting surface water, the
ClO2 concentrations are held at around 0.5–1.0 mg/l for continuous dosing, with the
actual value naturally depending on site conditions. If the dosing concentration ci,d,
ClO2 is reduced up to contact with the membrane by half to 0.3–0.5 mg ClO2/l, then,
according to Eq. (5.6), for the maximum allowable membrane contact time of
33,000–100,000 h and with continuous ClO2 dosing, there results a maximum
membrane service life τMOp max of 3.8–11.4 years.
For the use in practice of chlorine dioxide for seawater disinfection in SWRO
plants, the necessary condition is that appropriate ClO2 production processes are
selected that allow it to be generated with no chlorine content. Of the common
processes applied in water treatment for producing ClO2:

• chlorite/chlorine process
• chlorite acid process
• chlorite electrolysis process

the first named fails to meet this requirement, but the other two do. A detailed
description of the ClO2 manufacturing process with the reactions and other
particulars for the chemistry of the application of ClO2 for seawater desalination
are given under Sects. 2.2.1 and 2.2.1.1.2 in Volume 2.
The by-products of manufacturing chlorine dioxide solutions by means of the
chlorite acid process are chlorite ClO2 and chlorate ClO3. Both these compounds
also form when disproportioning ClO2 solutions. These have a strongly oxidizing
action and could therefore likewise damage the separation layer of polyamide
membranes. Lab tests with solutions of these compounds with a concentration of
100 mg/l ClO2 or ClO3 acting on seawater desalination membranes have shown,
though, that neither product flux nor salt rejection of the membranes are noticeably
reduced up to an exposure time of up to 400 h and also there are no detectable
changes in the chemical structure of the separation membrane layer [27].
For the above lab tests, most investigations were conducted with chlorine dioxide
solutions with the ClO2 present either in low-salinity water, sodium chloride solu-
tion, or synthetic seawater without supplementation with bromide. The operational
trials with chlorine dioxide were also undertaken with water that did not correspond
to the composition of seawater with regard to its bromide content. For this reason,
the influence of the bromide content specific for seawater was not assessed in these
investigations.
In fact, it had previously been assumed that chlorine dioxide could not oxidize the
bromide content of that seawater to bromine or to hypobromous acid. Nevertheless,
oxidation of bromide would appear to be possible if a large excess of it is present in
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 355

comparison to chlorine dioxide, for example, if ClO2 is used for online disinfection
of seawater in the feed to an SWRO plant. This is confirmed by the results of the
investigations described in [19] in which it was found with the Fujiwara test that after
just 16 h exposure of a seawater membrane to seawater containing bromide and
ClO2, halogenation of the membrane separation layer had taken place, and likewise
with x-ray spectroscopy, deposits of bromine in the membrane could be identified.
This means that halogenating bromine compounds had been formed by oxidation
reactions. If alongside oxidation, halogenation takes place, it may be expected that
the polyamide membrane will suffer greater damage. Under the conditions
prevailing for seawater, it may therefore be advisable, when estimating the maxi-
mum allowable membrane contact time τMOp max, to favour a lower value for
CTmembrane of around 10,000 mg ClO2/lh, or to dose chlorine dioxide intermittently,
like for chlorine disinfection, so as to correspondingly lengthen membrane service
lives τml max.

Impacts of Monochloramine and Tolerance Value


Chloramines are formed by the reaction of ammonia (NH3) or of ammonium
compounds (NH4Cl, (NH4)2SO4) with chlorine or with the hypochlorous acid
(HOCl) present in aqueous solution. Depending on the molar ratio of the two
reactants, in accordance with the following equations, monochloramine NH2Cl,
dichloramine NHCl2 or trichloramine NCl3, or a mixture of these compounds is
formed.

NH3ðaqÞ þ HOCl⇄NH2 Cl þ H2 O

NH2 Cl þ HOCl⇄NHCl2 þ H2 O

NHCl2 þ HOCl⇄NCl3 þ H2 O

However, the course of the reaction of chlorine with ammonia/ammonium is


highly complex, so the above equations are simplified representations.
Monochloramine is formed primarily within the pH range of 7 to 8.5. For pH values
of below 7, progressively more dichloramine is formed and then, below a pH of 4, to
an increasing degree trichloramine.
Of these chloramines, monochloramine possesses the most favourable
disinfecting and biocidal properties. This does less damage to membrane separation
layers than is the case with chlorine and chlorine dioxide. For this reason, at various
times consideration is given to the use of monochloramine as biocide and disinfec-
tant to suppress biological and organic fouling in SWRO plants, and especially,
those equipped with polyamide membranes [28, 29].
However, for direct contact with RO polyamide separation membranes, this
would require an almost pure substance, i.e. uncontaminated with the reactant
chlorine. Otherwise, like for chlorine dioxide, the polyamide separation membranes
could suffer increased damage from the combination of the two compounds. As
356 5 Reverse Osmosis Membrane System: Core Process of SWRO

stated by the membrane manufacturers, the content of free chlorine must be less than
0.1 mg Cl2/l.
For the formation of monochloramine, ammonium compounds (NH3, NH4Cl,
(NH4)2SO4) must be reacted with chlorine in a corresponding molar ratio. In addition
to the chemistry of chloramine formation, in seawater, the bromide it contains and
the hypobromous acid HOBr arising from the addition of chlorine react with the
ammonium compounds to form bromamines.

HOCl þ Br ⇄HOBr þ Cl

NH2 Cl þ HOBr⇄NHClBr þ H2 O

NHClBr þ HOBr⇄NHBr2 þ H2 O

Depending on the reaction conditions, i.e. the pH of the seawater and the ratio of
the concentration of chlorine to that of ammonium, various mixtures may be formed
of monochloramine and dichloramine, bromamines, or free bromine (the detailed
description of the chemical equilibria and the kinetics of monochloramine formation
under Sects. 2.2 and 2.2.1.1.3 in Volume 2).
Upon contact of monochloramine with polyamide membranes, their separation
layers are oxidized and also modified by halogenation [30]. Nevertheless, the
influence of NHCl2 on membrane performance is substantially less than is the case
for chlorine and chlorine dioxide. Manufacturers of polyamide membranes quote for
monochloramine in contact with their membranes values for CTmembrane in a range of
50,000–200,000 mg NH2Cl/lh, or even a value of up to 300,000 mg NH2Cl/lh. At
the same time, though, it is pointed out that these values apply only for
monochloramine solutions that contain virtually no chlorine. As is the case for the
other chlorine-containing disinfectants, oxidation and halogenation of polyamide
membrane separation layers by monochloramine are very strongly catalysed by
heavy metals, like Fe2+, Al+, Cu2+, so the time the biocide can act on the membranes
is also considerably reduced [31, 32]. Like also for chlorine and chlorine dioxide,
here too, the extent of membrane damage increases with temperature.
To ensure that primarily monochloramine is formed from the reactants ammonia
and chlorine and their compounds, and at the same time, to attain a very low chlorine
content in the reaction product, the reaction equilibrium must be shifted very close to
the monochloramine product. The necessary condition for this, as described under
Sects. 2.2 and 2.2.1.1.3 in Volume 2, is that the following reaction parameters must
be adjusted and monitored:

• an optimum reaction pH, pHoptim, that, depending on temperature, lies in a range of


7.7–8.3
• a ratio of the mole concentration of ammonia NH3 to that of chlorine Cl2 of
greater than 1
• a certain reaction time to allow equilibrium to be attained which in turn depends
on the reaction pH and on an excess of ammonia in the molar ratio NH3/Cl2:
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 357

Thus, the NH3/Cl2 molar ratio should not be less than 1.5 or, even better, it should
be 2.0 to avoid a rise of chlorine concentration in the permeate should there be a
deviation from the optimum reaction pH to a lower pH.

Because chlorine is a reactant in the equilibrium reaction, although its share in the
end product may be greatly reduced by appropriate reaction control, its presence
cannot be excluded.
If monochloramine is dosed at a concentration of 5 mg NH2Cl/l for disinfection
and if a contact concentration cØ,c,NH2Cl of 3 mg NH2Cl/l then remains for the
membrane separation layer, for CTmembrane values of 100,000–200,000 mg NH2Cl/
l, maximum allowable membrane contact times τc,NH2Cl max of 33,000–66,700 h are
calculated, or 3.8–7.6 years. However, whether such values for chloramination of
seawater are actually attainable is questionable.
In this case, due to the bromide content of the seawater, the reaction for chlora-
mine formation increases in complexity (Sects. 2.2 and 2.2.1.1.3 in Volume 2). This
also makes it more difficult to adjust the reaction conditions for seawater
chloramination and maintain these in practice so that the oxidizing and halogenating
influences on the membranes will still remain controllable. Up to now, there has been
no experience in practice of the use of chloramine in SWRO plants. For this reason,
to estimate the allowable membrane contact time τc, NH2Cl max, the lower range of
50,000–100,000 mg NH2Cl/lh should be favoured, in particular if there is a possi-
bility of catalytic action due to heavy metals and/or elevated temperatures. With a
value of 50,000 mg NH2Cl/lh and the above named contact concentrations of 3 mg
NH2Cl/lh, it would then still be possible to attain a value for τc, NH2Cl max of
16,700 h, equivalent to 1.9 years.
The chemical mechanisms for the direct action of inorganic chlorinated biocides
on polyamide membranes can be very complex, as shown above. Added to this are
the scarcely predictable influences of a catalytic nature and temperature fluctuations
on the intensity of the oxidation and halogenation processes in the separation
membrane layer. Regarding chlorine, direct loading of polyamide membranes at
the level of the contact concentrations that arise for disinfection may be excluded.
For chlorine dioxide and chloramine, too, with direct membrane contact, it is very
difficult to make reliable forecasts for the potential service life of the membranes.
Nearly all manufacturers of TFC polyamide membranes therefore stipulate that, if
chlorine and chlorinated disinfectants like chlorine dioxide and chloramine are used,
appropriate measures must be taken to remove these from the infeed to an RO plant
prior to its contact with their membranes such that the residual biocide in the form of
chlorine Cl2 is less than 0.1 mg/l (Sects. 5.5.2.4 and 5.5.2.4.2).
Only one membrane manufacturer—Hydranautics—does not completely forbid
the use of chlorine dioxide in direct contact with its membranes for purposes of
disinfection. The condition for this is that a ClO2 concentration of 1 mg/l not be
exceeded and the contact time will not amount to more than 1–4 h. Here, too, though,
it must once again be pointed out that the chlorine dioxide may not contain any free
chlorine and there may be no heavy metals, like Fe or Mn, present in the water or on
the membrane surfaces. Nevertheless, frequent membrane cleaning with ClO2 or
358 5 Reverse Osmosis Membrane System: Core Process of SWRO

daily dosing of it should not be done until more results from membrane tests or from
operational practice become available [33].

Influence of Acidic and Alkaline Feed Conditions


According to the manufacturers of spiral-wound elements with polyamide
membranes, a pH range of 2–11 is permissible during desalination operation of
their membrane elements. During chemical treatment of the membranes, i.e. for
short-term exposure, this pH range may even be extended to 1–12 or 1–13. This
means that, when undergoing chemical cleaning, this membrane type may be treated
with concentrated acids and alkalis. The allowable pH range applies for the mem-
brane elements as a whole, that is for the membranes with separation layer and
interlayers as well as for the materials used for the intake and permeate spacers and
the central permeate tube. As stipulated for the membrane elements from the same
manufacturer, also the materials of the pressure vessel in which the membrane
elements are installed must naturally comply with these pH requirements for opera-
tion and chemical cleaning of the RO module.

5.1.3.2.2 Chemical Impacts on Cellulose Acetate Membranes


Other than for polyamide membranes, the chemico-physical structure of cellulose
acetate (CA) membranes is modified with strongly acidic and alkaline pH values as
well as under oxidative conditions. To determine the allowable contact time of a CA
separation membrane, it is therefore necessary to consider both influencing factors
for which reason the pH of the solution to be desalinated as well as the concentration
of the oxidant it contains must be known.

Influence of Acidic and Alkaline Feed Conditions: Hydrolysis


With cellulose acetate membranes, the product flux and degree of salt rejection
depend greatly on the degree of acetylization of the cellulose substrate. With an
increasing proportion of acetylization groups, the salt rejection efficiency of a CA
membrane improves while its product flux drops [34].
CA membranes are subject to hydrolysis during which OH groups are formed
from the acetyl groups, CH3–CO–, of the cellulose acetate (Fig. 5.3); acetic acid
CH3COOH is formed from the split-off acetyl radical; and the degree of acetylation
of the membranes is reduced. Their transmission properties change correspondingly
and thus also their salt rejection capability. With a reduction of the degree of
acetylation, this rejection capability of the separation membranes drops but their
water permeability rises.
This process is greatly dependant on pH and temperature, with the hydrolysis rate
increasing substantially within the strongly acidic and alkaline pH ranges. A temper-
ature rise, too, accelerates the hydrolysis rate of the acetyl groups of the cellulose
acetate.
Investigations of the kinetics of cellulose acetate hydrolysis have shown that in
the temperature range of 10–40  C, the hydrolysis rate is at a minimum in the pH
range of between 4 and 5 (Fig. 5.16) [35]. On either side of a valley bottom pH range
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 359

Hydrolysis rate [sec-1]


1E-04

1E-05 Temperature [°C]


40
30
20
1E-06 10

1E-07

1E-08

1E-09

1E-10

1E-11
1 2 3 4 5 6 7 8 9 10 11 12
pH

Fig. 5.16 Cellulose acetate membranes: pH and temperature dependence of hydrolysis rate

of 3.5 to 6.0, the hydrolysis rate exhibits a rapid exponential rise in both the acidic
and the alkaline ranges.
When operating an RO plant equipped with CTA membranes, this means that the
pH of the inflow to the membranes should fluctuate within this minimum range to
avoid major changes in membrane performance. This applies also to in situ mem-
brane cleaning with chemicals, i.e. cleaning solutions should be used whose pH does
not significantly lie outside the valley bottom range of 3.5–6.0.

Chlorine: Impacts of Halogenation and Oxidation and Chlorine Tolerance


Values
Alongside hydrolysis, the structure of the CA membranes is influenced by oxidation
in that if their polymer structure is damaged, there will be consequential impacts on
their performance. The influence on water and salt transport characteristics is similar
to that of hydrolysis, i.e. both water permeability and salt transmission through the
membranes will increase. However, the influence of oxidation on the membrane
properties is not as great as is the case with hydrolysis. Further, CA and CTA
membranes are less sensitive to oxidation in comparison with polyamide
membranes. The kinetics of these two influencing factors—hydrolysis and oxida-
tion—then determines the overall kinetics of how the membrane changes and, in
particular, the change in the salt rejection capability of the CA/CTA membranes
[36]. For both oxidation and hydrolysis of the membranes, their water transmission
properties change much less than is the case for salt permeability.
360 5 Reverse Osmosis Membrane System: Core Process of SWRO

When defining a CTmembrane value, for CA/CTA membranes the pH must also be
stated for which this value applies. The reason for this is that the membrane structure
is influenced at the same time by both the pH through hydrolysis and the chlorine
concentration through oxidation [36]. When the salt transmission rate doubles, the
CTmembrane value of a CTA hollow-fibre membrane has a pH of 6.0 at 40,000 mg Cl2/
lh. If the pH increases to 6.5, the CTmembrane value, i.e. the membrane’s chlorine and
hydrolysis tolerance, drops to 15,000 mg Cl2/lh, that is by more than half. For a pH
of 6.0 and a chlorine contact concentration cØ,c,Cl2 of 0.5 mg Cl2/l, the membrane
exhibits a permissible contact time τc, Cl2 max of 80,000 h, or 9.1 years. At a pH of 6.5,
this permissible contact time is reduced to 30,000 h, or 3.4 years.
Under intermittent chlorination, CTA hollow-fibre membranes are exposed to
chlorine for 1 to 3 h per 24 h operating time. According to Eq. (5.8), this results in a
factor fct of 0.04 to 0.125 for the ratio of dosing time to the duration of the operating
period. By applying Eq. (5.7), for a permissible contact time τc, Cl2 max of 30,000 h at
a pH of 6.5, for the worst case, i.e. with an fct value of 0.125, a maximum membrane
lifetime τMOpmax of 27 years is calculated, referred to the hydrolysing and oxidation
actions. Consequently, under these operating conditions, the loss in performance of a
CTA membrane due to disinfection is negligible in comparison to other factors that
could influence the service life of RO membranes.

Chlorine Dioxide: Impact and Tolerance Value


In comparison with chlorine dioxide, a CTA RO membrane clearly has a still higher
oxidation stability than is the case for chlorine. When a CTA hollow-fibre module
was tested continuously for 6000 h with seawater of pH 6.5 and an average chlorine
dioxide concentration of 0.2 mg/l, there was virtually no change in the transmission
characteristics of its membranes regarding salt and water. The chlorine dioxide in
this case was generated by means of the acid/chlorite process [37].
Like for the polyamide membranes, for cellulose acetate membranes, too, the
extent and the rate of the change in their structure, and thus their water and salt
transmission characteristics, are catalysed by the chemical influence of heavy metals,
and these greatly speed up the process. For CTA membranes, this applies in
particular for the heavy metals copper, Cu, and cobalt, Co, as demonstrated by
investigations undertaken using Hollosep hollow-fibre membranes supplied by
Toyobo [38, 39].
However, the accelerating influence of catalysis by heavy metals on oxidative and
hydrolysing modifications to polyamide and CTA RO membranes can be lessened or
prevented by adding complexing agents. These compounds, also called chelating
agents, mask the heavy metals within a chemical structure that suppresses their
catalytic action. Such complexing agents are, for example, sodium
hexametaphosphate (SHMP) and organophosphonates, as used as antiscalants in
membrane desalination and also other organic complexing agents, like ethylenedi-
aminetetraacetate (EDTA) [38, 40].
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 361

5.1.4 Range of Operation Parameters and Conditions of Polyamide


and Cellulose Acetate Membranes

To match the range of mechanical and chemical stabilities specifically of the RO


membrane elements, there results certain conditions for operation of an RO mem-
brane desalination system that have to be maintained to ensure its reliability and a
correspondingly long membrane service lifetime. The primary operating parameters
that impose such limitations as specified by the membrane manufacturers are set out
in Table 5.2 for polyamide TFC spiral-wound elements as well as, by way of
comparison, for cellulose triacetate hollow-fibre membranes.

Table 5.2 Range of operation conditions of PA and CTA membranes


Type of membrane element
TFC polyamide spiral- HFF—cellulose
wound triacetate
Brackish Brackish
Parameter Unit Seawater water Seawater water
Feed pressure, pF max. bar ~83a ~41 ~67 to ~39
~82e
Static product backpressure, bar pF > pP; pPmax < 0.3–0.5b pF > pP; pPmax < 2
pPmax
Differential pressure, max.
• Of element 0.7–1.5b At max. Δp ¼ 1.0 bar,
• Of pressure vessel bar 2.0–4.0b chemical cleaning to be
started
Pressurization rate bar/min 30–35 20

Temperature max. C 45 40 35
pH range from . . . . to . . . .
max.
• During operation – 3–10.5; 2–11; 1–12c 3–8f
• During cleaning (short – 2–11; 1–11.5; 1–12; 1– 3–8
term) 12.5; 2–13; 1–13d
Chlorine contact mg/l Not detectable to <0.1 <1f
concentration, max.
Silt density index, SDI SDI15 5 4
(15 min), max.
a
Dependent on membrane manufacturer and feed temperature
b
Dependent on membrane manufacturer
c
Dependent on membrane manufacturer and feed temperature
d
Dependent on membrane manufacturer, cleaning temperature, and type of membrane
e
Dependent on feed temperature
f
Dependent on the quality and temperature of feed water
362 5 Reverse Osmosis Membrane System: Core Process of SWRO

Such operating limits are of a mechanical nature, for example:

• pressure applied to the membrane modules


• differential pressures arising at each of the membrane elements and the modules
• permissible pressure on the permeate side of the membrane elements
• limitations during pressure build-up at the membrane modules

as well as of a chemical or physical nature, such as:

• permissible pH range during desalination and chemical cleaning of the


membranes
• permissible maximum contact concentration for chlorine, chlorine compounds, or
free halogen calculated as free chlorine
• permissible particulate and colloidal load in the feed to the elements.

The permissible maximum operating pressure for RO membrane modules for


seawater desalination is greatly dependant on temperature (see Sect. 5.1.3.1.1). At
the maximum feed pressure to the membranes as quoted by the manufacturers, these
may only be operated at lower seawater temperatures, i.e. depending on membrane
type within a feed temperature range of around 10–20  C. At higher temperatures, the
permissible membrane pressure must be continuously reduced and, when seawater at
40  C is admitted to the membranes, this could drop by about 10–15 bar (Fig. 5.13).
When operating at high pressure and elevated temperature, the membranes are
compacted to an increasing degree. The membranes remain compacted even if
subsequently desalination continues at a lower temperature with the consequence
that, to obtain the same product flux as prior to the operation at elevated temperature,
the membrane feed pressure must be higher.
The other ranges for mechanical and chemico-physical operating limits as listed in
Table 5.2 differ depending on membrane manufacturer and specifically between
polyamide TFC and CTA-HFF membranes, as is to be expected. For the
manufacturers of polyamide membranes, the reasons for the differing values for the
operating parameters could be differences in the detailed membrane structure, but also
because, when specifying the parameters, the membrane manufacturers use different
preconditions. Thus, for example, the permissible differential pressure of a membrane
element or a membrane module is specified by one manufacturer on the basis of the
element’s mechanical stability, but by another under the aspect of limiting the build-
up of deposits on the membranes. A further factor is that membrane manufacturers
adopt in part differing safety philosophies for the permeate-side backpressure and also
for the permissible pH range or the use of oxidizing chemicals for chemical cleaning
of the membranes, or for the specific operating conditions of the RO system.
Due to the high packing density of the membrane elements (Table 5.6), these
have to be protected against the ingress of suspended and colloidal solids, as
otherwise, due to fouling in the flow channels and on the membrane surfaces, the
elements’ product recovery and salt rejection performance will be greatly reduced.
For this reason, the membrane manufacturers limit the fouling potential of the feed to
the elements by specifying maximum values for the silt density index (SDI). The
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 363

SDI is determined by application of a membrane filter approach for which the water
flow to be measured is admitted at a constant pressure of approximately 2 bar to a
membrane filter with a mean pore size of 0.45 μm. Then the time during which a
defined filtrate volume flows with an unclogged and with a clogged filter is measured
(for more details of the test method for determining the SDI, see Sects. 5.3 and
5.3.3.1). The manufacturers of spiral-wound module elements quote a maximum
value of 5 as silt density index SDI15 in the feed to their elements, whereas the
manufacturers of CTA-HFF modules quote a maximum value of 4 (Table 5.2).
However, pretreatment processes for the water to be desalinated should be so
selected that SDI values will be attained that are appreciably below the limit quoted
by the respective membrane manufacturer for the fouling potential. The higher the
SDI value in the feed to the membrane elements, the greater is the risk of the
membranes fouling, the more is the resulting decline in their performance, and the
greater is the frequency at which the membrane elements will have to be cleaned.

5.1.5 RO Membrane Design Basics

The basis for calculating and dimensioning an RO membrane desalination system is


knowledge of the rate of passage of water and salt through the semipermeable
separation membranes and how these are influenced by the operating conditions
during the desalination process, like salt content, temperature, and pressure of the
feed medium.
For derivation of the equations needed for this calculation, a number of models
are available; these differing in their methodology and in their understanding of the
nature of the passage of solution and solvent through the membrane separation layer
and the splitting of the two mass flows. The same applies for the assumptions for the
structure of the membrane separation layer, that is whether this is regarded as porous,
semi-porous, or homogeneous.
Basically, a differentiation is to be made between three types of models, these
being:

• diffusion-based models
• pore models
• models set up on the basis of irreversible thermodynamics.

Most of the media transport models for RO membranes can be assigned to one or
other of these main groups.
In the case of the diffusion-based models, the separation layer is assumed to be a
homogeneous pore-free medium with transport of salt and water taking place solely
due to diffusion through the membrane material. The separation action arises due to
the differing diffusion rates of the two media through the material of the separation
membrane. Volume flow and salt flow are mutually independent, i.e. they are not
interlinked. Counting to this type of model are, for example, the solution-diffusion
364 5 Reverse Osmosis Membrane System: Core Process of SWRO

model [41] and the solution-diffusion imperfection model [42]. As well as transmis-
sion due to diffusion, the latter also considers the influence of the pores.
For the pore model, it is assumed that the membranes contain numerous
micropores. Volumetric flow and salt flow are fixed by the differing sorption and
diffusion properties of the membrane material for solvent and dissolved substances
as well as the size of the pores and their structure. Also counting to this model type is
the preferential sorption-capillary flow model developed by Sourirajan in the course
of his research on cellulose acetate membranes [3].
For models based on irreversible thermodynamics, the transmission of salt and
that of water are considered to be linked and are derived solely on the basis of
thermodynamic principles, which are without considering the physical or chemical
properties of the separation membranes. Thermodynamic transport models are the
Kedem-Katchalsky Model [43] and the Spiegler-Kedem Model [44] that is a further
development of this.
Transport models that, alongside the mass passage due to sorption or diffusion,
consider additionally the surface charge of the separation membranes or the Donnan
potential that builds up on the surface of a charged separation membrane are of lesser
significance for mass passage processes in RO membranes that are used industrially
for seawater and brackish water desalination.
Despite their differing assumptions regarding the influence of the type of mem-
brane and its structure on the mechanism and extent of mass transport, all these
transport models mentioned above yield comparable results, this being that
generally:

• the volumetric flow through a separation membrane depends on how much the
difference of the pressure applied on the solution side exceeds the osmotic
pressure of the solution to be desalinated
• the salt flow through a separation membrane is determined by the difference
between the salt or component concentrations on the solution and permeate side.

5.1.5.1 Water and Salt Transport Through RO Membranes: Basic


Equations
The mass transport and concentration profiles that are present or build up at a
separation membrane are shown in Fig. 5.18. According to the mass flow parameters
and the concentrations on the solution side and the product side of the membrane,
under the influence of the pressure on the solution side, the concentration gradient on
the solution and product side and the individual transport properties of the separation
membrane, its specific volume, and salt transport properties are resulting. In particu-
lar, the degree of salt rejection of the separating membrane for the different
components of the solution to be desalinated determines the extent to which a
coupling of volume and salt transport through the membrane has to be considered
in the respective transport model. The algorithms derived from the Kedem-
Katchalsky/Spiegler-Kedem model for solvent and salt transport through a separa-
tion membrane include such a coupling of the passage of the solvent and the
components dissolved in it (Eqs. 5.9 and 5.10).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 365

The membrane flux JW, i.e. the water flow per surface unit of the separation
membrane, depends on a membrane-specific permeability coefficient A as well as on
the extent to which the difference between the solution- and the product-side
pressures ( pF  pP) exceeds the osmotic pressure difference between the solution
and the permeate sides, i.e. π M to π P (Eq. 5.9). However, the osmotic pressure
difference has still to be multiplied with a reflection coefficient σ, which takes into
account the degree of coupling between volumetric and salt flow.
The volumetric-related permeability coefficient A depends on the water perme-
ability KW of the membrane and the separation membrane’s thickness Δ. The
parameter KW in turn is influenced by the diffusivity of the water through the
membrane DMW and the temperature T (Eq. 5.9a).

J W ¼ A  ½pF  pP  σ  ðπ M  π P Þ: ð5:9Þ


DMW  cW  V W K W
A¼ ¼ ð5:9aÞ
RT Δ Δ
JW ¼ water (solvent) flux [kg/m2, h]
A ¼ water permeability coefficient of membrane [kg/m2hbar]
pF ¼ membrane feed pressure [bar]
pP ¼ product-side pressure [bar]
σ ¼ reflection coefficient
π M ¼ osmotic pressure at membrane wall [bar]
π P ¼ osmotic pressure of product [bar]
DMW ¼ diffusivity of water in membrane [m2/h]
cW ¼ water mean concentration in membrane [kg/m3]
VW ¼ partial molar volume of water 18.02  106 [m3/mol]
R ¼ universal gas constant ¼ 8.3144621  105 [m3bar/molK]
T ¼ temperature [K]
Δ ¼ membrane thickness [m]
KW ¼ water permeability of membrane [kg/mhbar]
The membrane flux of a component i of the salt content solution to be desalinated
is likewise dependent on a permeation coefficient BSi . This coefficient for the salt
transport, though, is specific not just for the separation membranes, but also for the
particular salt component i to be transported. The difference between the solution
side and permeate side concentrations of the salt components to be transported still
determines their rates of passage (Eq. 5.10). Added to this is another term in the
equation, which characterizes the degree of coupling between the volumetric and the
salt flow. This contains the membrane flux JW, the reflection coefficient σ, and the
logarithmic mean of the concentration ci,aver,M,P on the solution and the permeate
sides of the salt component concerned. Like for water transport, in the salt flux
algorithm too, the salt permeability coefficient BSi is dependent on the salt perme-
ability of the membrane K Si for the component i concerned and the thickness of the
separation membranes Δ. This parameter K Si is itself a function of the diffusivity of
the membrane DMSi for the component i to be separated, the thickness of the
366 5 Reverse Osmosis Membrane System: Core Process of SWRO

separation membranes Δ, and a distribution coefficient for the component i in the


membrane (Eq. 5.10a).
 
J Si ¼ BSi  ci,M  ci,p þ ð1  σ Þ  J W  ci,aver,M,P : ð5:10Þ

DMSi  k Si K Si
BSi ¼ ¼ : ð5:10aÞ
Δ Δ
J Si ¼ salt (solute) flux of membrane for component i [kg/m2, h]
BSi ¼ salt permeability coefficient of membrane for component i [m/h]
ci,M ¼ concentration of component i at membrane wall [kg/m3]
ci, P ¼ product-side concentration of component i [kg/m3]
ci,aver,M,P ¼ average membrane wall and product-side concentration of compo-
nent i [kg/m3]
DMSi ¼ diffusivity of component i in membrane [m2/h]
kSi ¼ distribution coefficient of component i in membrane []
K Si ¼ salt permeability of membrane for component i [m2/h]
For separation membranes for which their salt rejection factor Ri for the compo-
nent of the salt solution to be separated is in the vicinity of unity, the value of the
reflection coefficient σ can be set equal to 1, i.e. the degree of coupling of the
volumetric and salt transport is negligible [44]. For membranes for seawater and
brackish water desalination that are currently on the market with a salt rejection
factor under standard conditions in a range of 99.75–99.85% for seawater and up to
99.8% for brackish water, this assumption generally holds. For components whose
rejection efficiency is appreciably below these values, like for the separation of
boron for seawater desalination, for analysing salt rejection it would be appropriate
to apply a transport model that takes into account the coupling of volumetric and
salt flow.
If the reflection coefficient of the Spiegler-Kedem model takes a value of 1, then
the transport equations of this model are simplified to those of the solution-diffusion
model as given by Eqs. (5.11) and (5.12).

J W ¼ A  ½pF  pP  ðπ M  π P Þ: ð5:11Þ


 
J Si ¼ BSi  ci,M  ci,p : ð5:12Þ

Experience has shown that for membranes with high salt rejection, the mass
transport processes for ionogenic solution components can be calculated using the
solution-diffusion model with an outcome that is in good agreement with the results
of trials and operation of such membranes. The two transport equations for the
volumetric and salt flux of this model are therefore the basic algorithms from which
the follow-up equations are derived for dimensioning industrial RO systems for
seawater and brackish water desalination.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 367

5.1.5.1.1 Product (Solvent) Flux


The product flux rate JW of an RO separation membrane is dependent on the extent to
which the pressure difference between the solution and product sides of an RO
separation membrane pF  pP exceeds the osmotic pressure differential between the
solution to be treated and the permeate product π M  π P (Eq. 5.13). This driver for
the product flux is also termed the net driving pressure NDP (Eq. 5.13a).
  
J W ðtÞ ¼ AðtÞ  pF  pP  π M ðtÞ  π PðtÞ : ð5:13Þ
  
NDPðtÞ ¼ pF  pP  π M ðtÞ  π PðtÞ : ð5:13aÞ

JW(t) ¼ product flux at temperature t [kg/m2, h]


A(t) ¼ water (solvent) transport coefficient of membrane at temperature
t [kg/m2hbar]
π M(t) ¼ osmotic pressure at membrane wall at temperature t [bar]
π P(t) ¼ osmotic pressure of product at temperature t [bar]
NDP(t) ¼ net driving pressure at temperature t [bar]
But also the temperature of the solution to be treated governs the water transport
through the RO membrane. The reason for this is the dependency of both the water
permeability coefficient A and the osmotic pressure on temperature (Chap. 3 Sect. 3.
2.3.3, Eq. 3.115, Figs. 3.34 and 3.35).
If the water permeability coefficient At0 for a temperature t0 is known and should
the coefficient A(t) for a temperature t deviating from this have to be determined, At0
has be multiplied with a membrane-specific temperature coefficient fTCW
(Eq. 5.13b). The product flux at temperature t is then given by Eq. (5.13c).

AðtÞ ¼ At0  f TCW : ð5:13bÞ


  
J W ðt Þ ¼ At0  f TCW  pF  pP  π M ðtÞ  π PðtÞ : ð5:13cÞ

At0 ¼ water permeability coefficient of membrane at reference temperature t0


[kg/m2hbar]
fTCW ¼ temperature correction factor for water transport []
To calculate this temperature correction factor fTCW, normally the membrane
manufacturers provide an algorithm corresponding to Eq. (5.13d) or, for a reference
temperature of 25  C, in accordance with Eq. (5.13e). Depending on membrane type
and manufacturer, the coefficient kFW has a value in the range of 2500–3100. One
membrane manufacturer quotes differing values for the membrane factor kFW
depending on whether the temperature range is t  25  C or t  25  C.
h  i
AðtÞ k FW  273:15þt 0 273:15þt
1 1

f TCW ¼ ¼e ð5:13dÞ
At 0
368 5 Reverse Osmosis Membrane System: Core Process of SWRO

f TCW ¼ e½kFW ð298:15273:15þtÞ t 0 ¼ 25 C:


1 1
for ð5:13eÞ

kFW ¼ 2500  3100


kFW11 ¼ 2640 for t  25  C
¼ 3020 for t  25  C
kFW ¼ membrane factor []
t ¼ temperature t [ C]
t0 ¼ reference temperature [ C] ¼ 25  C
If for a certain membrane type, no temperature correction factors or equations for
calculating these are available, in ASTM standard D4516-00 “Standard Practice for
Standardizing Reverse Osmosis Performance Data”, application of Eq. (5.13f) is
proposed [45].

f TC ¼ f TCW ¼ f TCS ¼ 1:03t25 : ð5:13fÞ

In the graph shown in Fig. 5.17, the dashed curves show the temperature
correction factor fTCW plotted over the temperature range of 5–45  C, as calculated
using Eq. (5.13e) with a membrane coefficient kFW of 2700 and, by way of
comparison, the value of fTCW as calculated according to ASTM-D4516
(Eq. 5.13f). For this case, the two curves show good agreement.

Temperature correction
factor fTCW & fTCS
3.5

Temperature correction A
3.0
Temperature correction A ASTM D4516

2.5 Temperature correction Bi

2.0 kFW = 2,700


kFS = 5,000
1.5

1.0

0.5

0.0
0 10 20 30 40 50
Temperature [°C]

Fig. 5.17 Temperature correction factor for water and salt permeability coefficients A and Bi of RO
membranes

11
Datasource: DuPont Filmtec Reverse Osmosis Membranes Technical Manual Table 27.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 369

The water permeability of the membranes increases with rising temperature.


However, a rise in temperature results also in an increase of osmotic pressure
(Fig. 3.35), the consequence of which is that for a constant pressure pF and rising
temperature the value of the net driving pressure NDP is reduced. The positive
influence of the temperature rise on the water permeability is partially offset by the
temperature dependency of the osmotic pressure. This negative osmotic pressure
effect, though, is less than the influence of temperature on membrane permeability,
although it does increase with rising salt content and elevated temperature.
For calculating the membrane flux at various temperatures, apart from the tem-
perature dependency of water permeability, also the change in net driving pressure
NDP has to be considered. Upon combining Eq. (5.13c) for the product flux at
temperature t with Eq. (5.14) below for reference conditions, the relationship shown
in Eq. (5.15) is obtained from which the product flux JW(t) at a defined temperature
t can be calculated from a known product flux J Wt0 at a reference temperature t0.

J Wt0 ¼ At0  ½pF  pP  ðπ M,t0  π P,t0 Þ ¼ At0  NDPt0 : ð5:14Þ


f TCW  NDPðtÞ
J W ðtÞ ¼ J Wt0  : ð5:15Þ
NDPt0

J Wt0 ¼ water (solvent) flux at reference temperature [kg/m2, h]


NDPt0 ¼ net driving pressure at reference temperature t0 [bar]
π M,t0 ¼ osmotic pressure at membrane wall at reference temperature t0 [bar]
π P,t0 ¼ osmotic pressure of product at reference temperature t0 [bar]

5.1.5.1.2 Solute Flux and Salt Rejection


Solute Flux
The solute or salt flux J Si ðtÞ of an RO separation membrane for a component i is
calculated from the difference of its concentration on the solution side at the
membrane surface ci,M(t) and its concentration on the product side ci, p(t) (Eq. 5.16).
Like for the water permeability coefficient At, the salt permeability coefficient Bi(t) is
also temperature-dependent. Its value Bi(t) for a temperature t is derived from a
permeability coefficient Bi,t0 that is known for a reference temperature t0 by
multiplying this parameter with a temperature correction factor fTCS for the salt
transport (Eq. 5.16a). The salt flux at temperature t is then given by Eq. (5.16b).
 
J Si ðtÞ ¼ BiðtÞ  ci,M ðtÞ  ci,pðtÞ : ð5:16Þ

BiðtÞ ¼ Bi,t0  f TCS : ð5:16aÞ


 
J Si ðtÞ ¼ Bi,t0  f TCS  ci,M ðtÞ  ci,pðtÞ : ð5:16bÞ

J Si ðtÞ ¼ salt (solute) flux of component i at temperature t [kg/m2, h]


Bi(t) ¼ salt permeability coefficient of membrane for component i at temperature
t [m/h]
370 5 Reverse Osmosis Membrane System: Core Process of SWRO

Bi,t0 ¼ salt permeability coefficient of membrane for component i at reference


temperature t0 [m/h]
ci,M(t) ¼ concentration of component i at membrane wall at temperature t [kg/m3]
ci, p(t) ¼ product-side concentration of component i at temperature t [kg/m3]
fTCS ¼ temperature correction factor for salt (solute) transport []
The temperature correction factor for salt transport fTCS can be determined using
an algorithm that is similar to that for the temperature factor for water transport fTCW
(Eq. 5.16c).
h  i
BiðtÞ k FS  273:15þt 0 273:15þt
1 1

f TCS ¼ ¼e ð5:16cÞ
Bi,t0

fTCS ¼ temperature correction factor for salt transport []


The membrane manufacturers quote values for the membrane factor kFS ranging
from 5000 to 6000.
Figure 5.17 shows a graph of how the salt permeability coefficient Bi(t) varies with
temperature for a kFS value of 5000. The solid line shows the increase of Bi(t) with
rising temperature. Compared with the water permeability coefficient, A(t), it is seen
that the curve for the salt permeability coefficient, Bi, differs substantially from that
for A(t). Above 25  C, the increase of salt flux is markedly more pronounced than is
the case for the water permeability, i.e. the product flux. This means that the
temperature correction factor for salt transport fTCS cannot be set equal to the
temperature correction factor for water transport fTCW as stated according to
Eq. (5.13f) in ASTM standard D4516-00.
Combining Eqs. (5.16b) with (5.17) results in a relationship for calculating the
salt flux J Si ðtÞ at temperature t from the value J Si, t0 for a reference temperature t0
(Eq. 5.18).
 
J Si, t0 ¼ Bi,t0  ci,M,t0  ci,pðt0 Þ ð5:17Þ
 
f TCS  ci,M ðtÞ  ci,pðtÞ
J Si ðtÞ ¼ J Si, t0  ð5:18Þ
ci,M,t0  ci,pðt0 Þ

J Si t0 ¼ salt (solute) flux of component i at reference temperature t0 [kg/m2, h]


ci,M,t0 ¼ concentration of component i at membrane wall at reference temperature
t0 [kg/m3]
ci,pðt0 Þ ¼ product-side concentration of component i at reference temperature t0
[kg/m3]
Because a change in temperature changes the salt flow through the membrane and
thus at the same time the product concentration cip and consequently also the
concentration of the component i at the membrane wall ci,M, Eq. (5.18) must be
solved iteratively. For an RO membrane for seawater desalination, however, the
product concentration is very low compared to the solution concentration
(ci, p ci, M), which means that the solution concentration, too, can be set equal
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 371

at a temperature t to the concentration at t0 (ci,M,t0 ¼ ci,M ðtÞ Þ. Equation (5.18) can then
be simplified as shown by Eq. (5.19).

J Si ðtÞ ¼ f TCS  J Si, t0 : ð5:19Þ

Solutes Rejection
The concentration of a salt component ci, p(t) in the product solution generated by the
separation process results from a mass balance for the salt and product flux at the
separation membrane. ci, p(t) is calculated from the ratio of the salt flux J Si ðtÞ of the
component i to the product flux JW(t) of the separation membrane or as the quotient of
the salt flux and the sum of both flux rates (Eqs. 5.20 and 5.20a).

J Si ðt Þ
ci,pðtÞ,m ¼ : ð5:20Þ
J Si ðt Þ þ J W ðt Þ

J Si ðtÞ  ρpðtÞ
ci,pðtÞ,v ¼ ci,pðtÞ ¼ : ð5:20aÞ
J Si ðtÞ þ J W ðtÞ

ci, p(t), m ¼ product-side concentration of component i at temperature t [kg/kg]


ci, p(t), v ¼ product-side concentration of component i at temperature t [kg/m3]
ρp(t) ¼ density of product solution [kg/m3]
For the seawater desalination separation process, salt flux is very low compared to
product flux (J Si ðtÞ J W ðtÞ ). Further, due to the low product-side salt content, the
product solution’s density can be set equal to that of pure water (ρp(t) ρW(t)).
Equation (5.20a) is then simplified to Eq. (5.20b). The density of water ρW(t) as a
function of temperature may be calculated using Eq. (5.20c).

J Si ðtÞ  ρW ðtÞ
ci,pðtÞ ¼ : ð5:20bÞ
J W ðt Þ

ρW ðt Þ ¼ a1 þ a2  t þ a3  t 2 þ a4  t 3 þ a5  t 4 : ð5:20cÞ

a1 ¼ +9.9992  102
a2 ¼ +2.0341  102
a3 ¼ 6.1625  103
a4 ¼ +2.2615  105
a5 ¼ 4.6571  108
ρW(t) ¼ density of water [kg/m3]

By dividing the mass product flux JW(t) in [kg/m2, h] by the density of water ρW(t)
in [kg/m3], the volumetric product flux JW(t),V in [m3/m2, h], that is more commonly
used for technical calculations of membrane plants, is obtained (Eq. 5.20d) and the
concentration of the component i in the product is then found using Eq. (5.20e).
372 5 Reverse Osmosis Membrane System: Core Process of SWRO

J W ðtÞ
J W ðtÞ,V ¼ : ð5:20dÞ
ρW ðtÞ

JW(t), V ¼ product flux at temperature t [m3/m2, h]

J Si ðtÞ
ci,pðtÞ ¼ : ð5:20eÞ
J W ðtÞ,V

For calculating the volumetric-based product flux JW(t),V and then also the volu-
metric water transport coefficient A(t),V (Eq. 5.20g), Eq. (5.13) is transformed into the
form shown by Eq. (5.20f).
  
J W ðtÞ,V ¼ AðtÞ,V  pF  pP  π M,ðtÞ  π P,ðtÞ : ð5:20fÞ

AðtÞ
AðtÞ,V ¼ : ð5:20gÞ
ρW ð t Þ

A(t), v ¼ volumetric-based water (solvent) transport coefficient of membrane at


temperature t [m3/m2hbar]
If Eq. (5.20e) is combined with Eqs. (5.20f) and (5.16) for the salt flux of
component i J Si ðtÞ , the product salt content ci,p(t) of component i is obtained as
functions of its salt permeability Bi(t), the water permeability of the separation
membrane A(t),V, the concentration gradient of component i at the membrane ci, M(-
t)  ci, p(t), and the net driving pressure NDP(t), pF  pP  (π M(t)  π P(t)) (Eq. 5.21).

 
BiðtÞ  ci,M ðtÞ  ci,pðtÞ
ci,pðtÞ ¼    : ð5:21Þ
AðtÞ,V  pF  pP  π M ðtÞ  π PðtÞ

If the two flux rates are combined to a permeability quotient Pi(t) according to
Eq. (5.21a) and this, together with the net driving pressure NDP(t) substituted into
Eq. (5.21), Eq. (5.21b) is obtained.

AðtÞ,V
PiðtÞ ¼ : ð5:21aÞ
Bi,ðtÞ
ci,M ðtÞ ci,M ðtÞ
ci,pðtÞ ¼ : ð5:21bÞ
PiðtÞ  NDPðtÞ þ 1 PiðtÞ  NDPðtÞ

Pi(t) ¼ permeation quotient at temperature t [bar1]


The more the permeability quotient, Pi(t) increases, the lower the salt transport
through a membrane is in proportion to its product flux. This therefore provides a
performance indicator for a separation membrane regarding production performance
and salt rejection. As shown by Eq. (5.21b), the concentration of a salt component
i in the product decreases with increasing permeability quotient Pi(t) and net driving
pressure NDP(t).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 373

The intrinsic salt passage of an RO separation membrane for a component i is


determined by its concentration in the product cip(t) and at the membrane wall ci,M(t)
(Eq. 5.22).

ci,pðtÞ
SPi ðtÞ ¼ : ð5:22Þ
ci,M ðtÞ

If Eq. (5.22) for the salt passage factor SPi(t) is combined with Eq. (5.21) to
calculate the product concentration of the component i ci,p(t), there results a relation-
ship for SPi(t), that likewise shows its dependency on the membrane permeabilities
A(t),V and Bi(t) or the permeability quotient Pi(t) and the net driving pressure NDP(t)
(Eqs. 5.23 and 5.24). Here, too, it can again be seen that the salt passage of a
separation membrane decreases as the permeability quotient and the net driving
pressure increase.

    1
AðtÞ,V  pF  pP  π M ðtÞ  π PðtÞ
SPi ðtÞ ¼ 1þ : ð5:23Þ
BiðtÞ

1  1
SPi ðtÞ ¼ ¼ 1 þ PiðtÞ  NDPðtÞ : ð5:24Þ
1 þ PiðtÞ  NDPðtÞ

By applying the relationship for the salt rejection factor Ri(t) as shown in
Eq. (5.25) then, like for the salt passage factor SPi(t), equations for salt rejection
can be derived (Eqs. 5.26 and 5.27).

RiðtÞ ¼ 1  SPi ðtÞ : ð5:25Þ

1  1
RiðtÞ ¼ 1  ¼ 1  1 þ PiðtÞ  NDPðtÞ : ð5:26Þ
1 þ PiðtÞ  NDPðtÞ
 1  1
BiðtÞ 1
RiðtÞ ¼ 1þ ¼ 1þ ð5:27Þ
AðtÞ,V  NDPðtÞ PiðtÞ  NDPðtÞ

Ri(t) ¼ membrane (intrinsic) salt rejection factor of component i at temperature


t []

5.1.5.1.3 Concentration Polarization and Membrane Performance


Concentration Polarization
If a feed solution flow FF sweeps along an RO separation membrane under pressure,
then a concentration profile builds up between a Point X in the direction of flow and
an associated Point Y on the membrane surface that depends primarily on the flow
conditions between points X and Y and the volumetric flow of the desalinated
product Fp through the membrane at Point Y. The concentration of a solution
component i increases from its original concentration in the feed flow ci,F to the
so-called bulk concentration ci,b at Point X in line with the water flow Fp that passes
374 5 Reverse Osmosis Membrane System: Core Process of SWRO

through the membrane as desalinated product. At the membrane itself, the concen-
tration increases even more up to the membrane wall concentration ci,M due to the
substantially higher flow of product water Fp through the membrane in comparison
with the salt flow FS (Fig. 5.18).
The variation of this component concentration ci within this concentration profile
with its dependency on the flow conditions on the solution side and on water flow on
the membrane’s product side can be calculated by applying thin film theory.
Accordingly, at the membrane’s surface on the feed side, a solution boundary
layer is formed within which essentially concentration increases due to the separa-
tion of the water from the salt at the membrane. The rise in concentration at this
solution-side boundary layer is termed concentration polarization. This concentra-
tion polarization determines the salt and component concentration ci,M at the mem-
brane surface and thus also the concentration ci,p in the desalinated product water.
The density δ of the solution boundary layer is a function of the ratio of the
product flow Fp through the membrane and the bulk feed flow FF of the solution with
the resulting flow profile. Likewise dependent on the latter is the share of the
component i that is transported to the membrane wall, is there retained in the
boundary layer or, due to diffusion backflow FD and backmixing, is returned to
the solution.
The relationship between the various component concentrations, the product flux
of the separation membrane as well as the diffusion properties of component i in the
solution and in water plus the thickness of the solution boundary layer is given by
Eq. (5.28).

Concentration in product ci,P

Product
Product flow F P

Salt flow FS
RO Membrane
Concentration at membrane ci,M
Boundary layer δ
Diffusion back flow FD

Feed flow FF Y
Bulk concentration c i,B

Feed / bulk solution

Concentration in feed ci,F

Fig. 5.18 RO membrane: flow and concentration profiles


5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 375

ci,M  ci,pðtÞ J W ðtÞ  δB


ln ¼ : ð5:28Þ
ci,B  ci,pðtÞ DSiðtÞ

δB ¼ thickness of boundary layer [m]


DSi(t) ¼ diffusion coefficient of component i in solution at temperature t [m2/h],
[m2/s]
kCP(t) ¼ concentration polarization mass transfer coefficient at temperature
t [m/h], [m/s]
The ratio of the solution-side diffusion coefficient of component i DSi(t) to the
thickness of the solution boundary layer δB is termed the concentration polarization
mass transfer coefficient kCP(t) (Eq. 5.28a). Resulting from Eqs. (5.28) and (5.28a) is
the exponential function Eq. (5.29) describing the concentration polarization at an
RO membrane.

DSiðtÞ
¼ k CPðtÞ : ð5:28aÞ
δB
J W ðt Þ
ci,M ðtÞ  ci,pðtÞ
¼ ekCPðtÞ ¼ M PðtÞ : ð5:29Þ
ci,B  ci,pðtÞ

MP(t) ¼ polarization modulus []


The concentration polarization factor βi(t) is a measure of the ratio of the concen-
tration of a component i at the membrane wall ci,M to its bulk concentration ci,B
(Eq. 5.30).

ci,M ðtÞ
βiðtÞ ¼ : ð5:30Þ
ci,B

βi(t) ¼ concentration polarization factor at temperature t []


Its value provides the basis for calculating the salt or component concentration at
the membrane wall ci,M that determines the salt flow through the separation mem-
brane and, in turn, the osmotic pressure critical for the membrane’s product flux π M(t)
that results from these ci,M values.
Due to the differing salt permeability coefficients Bi(t) and thus the differing
rejection rates Ri(t) of the solution components by the membrane, the concentrations
of the various components likewise differ in the solution boundary layer. This effect
can be taken into account by combining Eq. (5.30) with Eq. (5.31) for the rejection
factor Ri(t). The concentration polarization factor β(t) can be calculated with the
resulting Eq. (5.32), also under consideration of the specific rejection factors Ri(t)
of a membrane.

ci,M ðtÞ  ci,pðtÞ ci,pðtÞ


RiðtÞ ¼ ¼1 ð5:31Þ
ci,M ðtÞ ci,M ðtÞ
376 5 Reverse Osmosis Membrane System: Core Process of SWRO

J W ðt Þ

ekCPðtÞ M PðtÞ
βiðtÞ ¼ ¼   : ð5:32Þ
  J W ðt Þ
RiðtÞ þ 1  RiðtÞ  M PðtÞ
RiðtÞ þ 1  RiðtÞ  e k CPðtÞ

Assuming the rejection factor of each salt component is around unity (Ri(t) ≌ 1),
which is permissible for the principal components of the salt content of seawater, the
equation is simplified to Eq. (5.32a).
J W ðt Þ

βðtÞ ffi ekCPðtÞ ffi M P : ð5:32aÞ

The mass transfer coefficient of the concentration polarization kCP(t) is fixed to an


overwhelming degree by the geometry of the membrane-side flow cross-section. In
practice, this is determined empirically for each membrane element type and its
specific membrane configuration and design. For a particular solution component i,
this can be done on the basis of Eq. (5.29) under defined flow conditions, such as
concentration, temperature, and product flux.
The values for kCP(t) under differing operating and flow conditions may be
calculated by means of similarity relationships as are applied for calculating convec-
tive mass and heat transfer. Thereby a Sherwood correlation according to Eq. (5.33)
finds application in which the Sherwood number Sh is expressed as a function of the
Reynolds and Schmidt numbers Re and Sc.

k CPðtÞ  dH
Sh ¼ ¼ a  Re b  Scc : ð5:33Þ
DSiðtÞ

Sh  DSiðtÞ DSiðtÞ
kCPðtÞ ¼ ¼ a  Re b  Scc  : ð5:33aÞ
dH dH
ρFðtÞ  νeff  dH
Re ¼ : ð5:33bÞ
μF ðt Þ
μF ðtÞ
Sc ¼ : ð5:33cÞ
ρF ðtÞ  DSiðtÞ

Sh ¼ Sherwood number []


Re ¼ Reynolds number []
Sc ¼ Schmidt number []
νeff ¼ effective velocity in flow channel [m/s]
ρF(t) ¼ feed water density at temperature t [kg/m3]
μF(t) ¼ feed water dynamic viscosity at temperature t [kg/ms]
dH ¼ hydraulic diameter of flow channel [m]
a, b, c ¼ empirical values that depend on flow conditions, type, and
characteristics of element
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 377

Element type a b c References


Hollow-fine fibre CTA 0.048 0.600 0.333 [46]
Spiral-wound TFC 0.080 0.875 0.250 [47]
0.065 [48]

The coefficient a as well as the exponents b and c in the Sherwood correlation


Eq. (5.33) are membrane element-specific parameters. The mass transfer coefficient
kCP(t), the Reynolds number Re, and the Schmidt number Sc are calculated as shown
respectively in Eqs. (5.33a), (5.33b), and (5.33c). Values for the diffusion coefficient
DSi(t) for the principal ions in seawater with relationships for its dependence on
salinity and temperature are to be found in [49], calculation formulas and numerical
values for the density of the feed water to be desalinated ρF(t) in Chap. 3, Sect. 3.2.
2.1, Eqs. (3.39), (3.40) and Fig. 3.15, and for its dynamic viscosity μF(t) in Chap. 3,
Sect. 3.2.2.2, Eqs. (3.41), (3.42) and Fig. 3.16. For spiral-wound modules,
calculations of the velocity νeff and the hydraulic diameter dH are described under
Sect. 5.1.5.2.3 with the associated algorithms Eqs. (5.59c)–(5.59e) and (5.59f)–
(5.59i).
The concentration polarization factor β(t) is determined by Eqs. (5.32) and (5.32a)
through the product flux JW(t) of the membrane and the parameters of temperature
plus net driving pressure NDP that influence this in accordance with Eq. (5.41) as
well as by the mass transport coefficient kCP(t).
The mass transport coefficient kCP(t) is characterized by the flow conditions
directly at the membrane, particularly whether flow is laminar or turbulent, and in
accordance with the Sherwood correlation (Eq. 5.33) with calculation of its
associated parameters Reynolds number Re (Eq. 5.33b) and Schmidt number Sc
(Eq. 5.33c). It is also influenced by both the effective flow velocity veff and the
density ρF(t) as well as the viscosity μF(t) with its dependency on temperature and the
salt content of the solution to be treated. This means that measures for improving
mixing and backmixing within the membrane feed flow cross-section have a sub-
stantial influence on the mass transport coefficient kCP(t). For spiral-wound modules,
important in this regard are in particular the height of the feed channel and the design
of the spacers mounted in it, their thickness, and their geometry.
Concentration polarization is a parameter that not only critically influences
production performance and the separation behaviour of an RO membrane, but—
as will be shown later—it is also important for the scaling and fouling characteristics
of a membrane desalination plant. When designing membrane elements and
arranging a number of these in modules, particular attention has therefore to be
paid to minimizing the concentration polarization. In commercial plants, according
to the stipulations of the membrane manufacturers, the concentration polarization
factor β(t) should not exceed 1.2. Usually, however, the plants are operated at a value
for β(t) of around 1.1 and less. For compliance with these values, it is important to set
limits to the ratio of membrane flux to throughflow within the membrane elements
and to always remain within these during operation.
378 5 Reverse Osmosis Membrane System: Core Process of SWRO

Membrane Performance
The concentration of a component i of the salt content of the water to be treated at the
surface of a separation membrane ci,M or rather the osmotic pressure π M(t) resulting
from this component concentration is decisive for the product flux (Eq. 5.13) of the
membrane and for the salt flux of component i (Eq. 5.16).
The membrane concentration of component i ci, M(t) is calculated by multiplying
its bulk concentration ci, B(t) with the concentration polarization factor βi(t) (Eq. 5.34),
and in turn the bulk concentration ci, B(t) is derived from the initial concentration of
component i ci,F in the water to be treated by multiplying ci,F with a concentration
factor CFi (Eq. 5.35).

ci,M ðtÞ ¼ βiðtÞ  ci,BðtÞ : ð5:34Þ

ci,BðtÞ ¼ ci,F  CFi : ð5:35Þ

ci,B(t) ¼ bulk concentration of component i at temperature t [kg/m3] [g/l]


CFi ¼ concentration factor of component i []
The concentration factor CFi depends on the product recovery factor Y and the
factor for the membrane’s salt passage for the component i SPi(t), or alternatively, the
corresponding factor for salt rejection Ri(t) (Eqs. 5.36 and 5.36a). If the membrane’s
salt rejection is almost 100% (Ri(t) ≌ 1), the known equation for the concentration
factor is obtained in accordance with Eq. (5.36b).
 
1  Y  SPi ðtÞ 1  Y  1  RiðtÞ
CFi ¼ ¼ : ð5:36Þ
1Y 1Y
CFi ¼ ð1  Y ÞRi : ð5:36aÞ
1
CFY ¼ : ð5:36bÞ
1Y
Y ¼ recovery coefficient
CFY ¼ concentration factor based on recovery Y only []
From the initial concentration of the component ci,F of the salt content of the
water to be treated, the concentration factor CFi, and the concentration polarization
factor βi(t), there then results in accordance with Eq. (5.37) the concentration at the
membrane wall ci,M(t) that is the determining factor for the performance of an RO
separation membrane.

ci,M ðtÞ ¼ ci,F  CFi  βiðtÞ : ð5:37Þ

Similar to the membrane component concentration ci,M(t), the solution-side


osmotic pressure at the membrane wall π M(t) can be determined from the osmotic
pressure of the initial solution π F(t) (Eq. 5.38).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 379

π M ðtÞ ¼ π FðtÞ  CF  βðtÞ : ð5:38Þ

However, this can also be calculated from the membrane component


concentrations of the solution boundary layer if these have all previously been
determined in accordance with Eq. (5.37). As shown in Eq. (5.39), the product-
side osmotic pressure π P(t) may be estimated.
 
π PðtÞ ffi π M ðtÞ  1  RSðtÞ ffi π M ðtÞ  SPSðtÞ : ð5:39Þ

RSðtÞ ¼ rejection factor of membrane for salt content at temperature t []


SPSðtÞ ¼ salt passage factor of membrane for salt content at temperature t []
Naturally, the calculation will be more accurate if it is done by taking the actual
ionogenic composition of the product.
Calculation of the osmotic pressure is described in Chap. 3, Sect. 3.2.3.3. Here the
equations needed for performing this calculation are given (Eqs. 3.115, 3.115a–3.
115g) together with algorithms for determining the osmotic coefficient (Eqs. 3.116a,
3.117, and 3.118). Values for the osmotic coefficient and the osmotic pressure of
seawater as functions of salinity and temperature are shown there in Figs. 3.34 and
3.35.
Also the ASTM standard “ASTM D4516—Standard Practice for Standardizing
Reverse Osmosis Performance Data” includes an equation for calculating osmotic
pressure from the molality of the salt components (Eqs. 5.40 and 5.40a).
X
π FðtÞ ¼ 8:308  ∅ðtÞ  ð273:15 þ t F Þ  mmi,F ½kPa: ð5:40Þ
i
X
π F ðtÞ ¼ 8:308  102  ∅ðtÞ  ð273:15 þ t F Þ  mmi,F ½bar: ð5:40aÞ
i

mmi,F ¼ molality of component i [mol/kgH2O]


For calculating the water and salt fluxes from the salt content of the initial solution
and the concentrations of the salt content components, the following Eqs. (5.41) and
(5.42) are derived from the basis equations for these membrane parameters
(Eqs. 5.13 and 5.16) together with Eqs. (5.37) and (5.38). The volumetric product
flux JW(t),V is calculated using Eq. (5.41a). The net driving pressure NDP(t) of the
product flux equation is determined by Eq. (5.41b).
h  i
J W ðtÞ ¼ AðtÞ  pF  pP  βðtÞ  π FðtÞ  CFY  π PðtÞ : ð5:41Þ
h  i
J W ðtÞ,V ¼ AðtÞ,V  pF  pP  βðtÞ  π F ðtÞ  CFY  π PðtÞ : ð5:41aÞ
h  i
NDPðtÞ ¼ pF  pP  π FðtÞ  CFY  βðtÞ  π PðtÞ : ð5:41bÞ
380 5 Reverse Osmosis Membrane System: Core Process of SWRO

 
J SiðtÞ ¼ Bi,ðtÞ  ci,F  CFi  βiðtÞ  ci,pðtÞ : ð5:42Þ

Equation (5.21b) for calculating the component concentration ci,p(t) in the product
and the product salt content cp(t) is transformed by application of Eqs. (5.41a),
(5.41b), and (5.42–5.43) and then to Eq. (5.43a).

ci,F  CFi  βiðtÞ


ci,pðtÞ ¼ h  i : ð5:43Þ
PiðtÞ  pF  pP  π F ðtÞ  CFy  βiðtÞ  π PðtÞ þ 1

ci,F  CFi  βðtÞ ci,F  CFi  βðtÞ


ci,pðtÞ ¼ : ð5:43aÞ
PiðtÞ  NDPðtÞ þ 1 PiðtÞ  NDPðtÞ

For determining the salt passage factor SPi(t) (Eq. 5.24) and the salt rejection
factor Ri(t) (Eq. 5.27), the net driving pressure NDP(t) must be calculated using
Eq. (5.41b).
If the membrane-specific salt passage factor SPi(t) or the salt rejection factor Ri(t) is
known, the concentration of the component i in the product is calculated by
combining Eqs. (5.22) and (5.37) which then results in Eq. (5.44).
 
ci,pðtÞ ¼ ci,F  CFi  βðtÞ  SPi ðtÞ ¼ ci,F  CFi  βðtÞ  1  RiðtÞ : ð5:44Þ

If the salt passage factor SP is referred to solution-side component concentrations


other than the concentration at the membrane wall ci,M(t) (Eq. 5.45), for example the
bulk concentration ciB(t) (Eq. 5.45a) or the feed concentration ciF (Eq. 5.45b), then
the membrane-specific salt passage factor SPi(t) is calculated from these salt passage
factors SPiB ðtÞ and SPiF(t) as shown respectively in Eqs. (5.46) and (5.47).

ci PðtÞ
SPiM ðtÞ ¼ : ð5:45Þ
ci,M ðtÞ
ci PðtÞ,
SPiB ðtÞ ¼ : ð5:45aÞ
ci BðtÞ
ci PðtÞ
SPiF ðtÞ ¼ : ð5:45bÞ
ci F ð t Þ

SPiB ðtÞ
SPiM ðtÞ ¼ : ð5:46Þ
βiðtÞ

SPiF ðtÞ
SPiM ðtÞ ¼ : ð5:47Þ
CFi  βiðtÞ

SPiM(t) ¼ specific membrane salt passage factor of component i based on bound-


ary layer concentration ¼ intrinsic salt passage of component i []
SPiB(t) ¼ salt passage factor of component i based on bulk concentration []
SPiF(t) ¼ salt passage factor of component i based on feed concentration []
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 381

5.1.5.2 Membrane Element Calculations


The basic equations given above under Sects. 5.1.5.1.2–5.1.5.1.3 for the mass flow
and salt rejection of an RO desalination membrane describe these processes only
considering selected points, that is, just for the x, y coordinates of a specific point on
a membrane. Within a membrane element, this separation membrane is swept by the
water to be treated not only at certain points but across its entire surface, with the
resulting product water also exiting across the surface on its opposite side. On both
the feed and product sides, depending on the type of element, concentration and flow
conditions become established specific for the membrane and the element which
have to be factored into the calculation (Fig. 5.18). Further, chemical and ageing-
related influences that change the separation membrane’s permeability and its salt
rejection behaviour as the period of service increases have to be taken into account.
For calculating the performance of an industrial membrane element, the basic
algorithms must therefore be adapted and extended to match the conditions encoun-
tered in practice.
This concerns the following parameters:

• average feed-to-brine bulk concentration for calculating water and salt flows
• concentration polarization
• pressure loss at the membrane element.
• long-term operation correction factor for the membrane element’s reduction of
permeability and salt rejection over the membrane’s service life.

5.1.5.2.1 Average Feed-to-Brine Bulk Concentration and Osmotic Pressure


The salt content on the feed side of a membrane element increases as the mass flows
of salt and water through the membranes change. At the same time, the salt content in
the product also increases as it is dependent on the feed side salt content. This salt
content results from mixing the differing product salt contents across the
membrane’s surface. Because, as the basic equations show, the product flux of the
membranes and the salt passage into the product depend on the local solution-side
concentration, for calculating the product composition for a membrane element,
instead of a feed-side concentration at a certain point at the membrane wall ci, M(t); an
average concentration value ciFC,M,EðtÞ has to be determined that as closely as
possible approximates the change in concentration between feed and concentrate
discharge (Fig. 5.18). Like for the point-by-point approach, this average membrane
wall concentration is calculated from the concentration of the feed to the membrane
element ci, F, E, but with an average value for the concentration factor CFiE and
likewise an average value for the concentration polarization factor βiEðtÞ (Eq. 5.48).
The average of the concentration polarization factor results from the quotient of the
average concentration of the component i or the salt content at the membrane wall
and its average bulk concentration (Eqs. 5.48a–5.48d).
382 5 Reverse Osmosis Membrane System: Core Process of SWRO

ciFC,M,EðtÞ ¼ ciFC,B,EðtÞ  βiEðtÞ ¼ ciFE  CFiE  βiEðtÞ : ð5:48Þ

ciFC,M,EðtÞ
βiEðtÞ ¼ : ð5:48aÞ
ciFC,B,EðtÞ

csFC,M,EðtÞ
βsEðtÞ ¼ : ð5:48bÞ
csFC,B,EðtÞ
X
csFC,M,EðtÞ ¼ ciFC,M,EðtÞ : ð5:48cÞ
i
X
csFC,B,EðtÞ ¼ csFC,B,EðtÞ : ð5:48dÞ
i

CFiE ¼ average concentration factor of element for component i []


βiEðtÞ ¼ average concentration polarization factor of component i at temperature
t []
βsEðtÞ ¼ average concentration polarization factor of salt content at temperature
t []
ciFE ¼ element feed concentration of component i [kg/m3] [g/l] [mg/l]
ciFC,M,EðtÞ ¼ average feed-to-concentrate concentration of component i at the
element membrane wall at temperature t [kg/m3] [g/l] [mg/l]
ciFC,B,EðtÞ ¼ average feed-to-concentrate bulk concentration of component i of
element at temperature t [kg/m3] [g/l] [mg/l]
csFC,M,EðtÞ ¼ average feed-to-concentrate concentration of salt at the element
membrane wall at temperature t [kg/m3] [g/l] [mg/l]
csFC,B,EðtÞ ¼ average feed-to-concentrate bulk concentration of salt at temperature
t [kg/m3] [g/l] [mg/l]
The average of the concentrations between feed and discharge of a membrane
element, i.e. the average bulk concentration of the element ciFC,B,EðtÞ , is the average
of the concentrations across all points of the membrane surface as calculated with the
concentration factor CFi according to Eq. (5.36a). An algorithm for this concentra-
tion average or actually for the concentration factor CFiEðintÞ with which the feed
concentration to the element ciFE has to be multiplied to calculate this value
(Eq. 5.49) is consequently obtained by differentiating Eq. (5.36a) and integrating it
within the limits 0 to YE [50].

1  ð1  Y E Þð1RiEðtÞ Þ
CFiEðintÞ ¼   : ð5:49Þ
Y E  1  RiEðtÞ

CFiEðintÞ ¼ average concentration factor of element integrated from YE ¼ 0 > x []


If Ri(t) ≌ 1, that is with Eq. (5.36b, in the same way as described above it is
possible to derive the known logarithmic relationship for forming the average from
the feed and discharge concentrations of the membrane element (Eq. 5.50).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 383

h i
1
ln 1Y E
CFiEðlÞ ¼ : ð5:50Þ
YE
CFiEðlÞ ¼ average concentration factor of element (logarithmic average) []
For separation membranes with high rejection efficiency (R  0.99) as is the case
for seawater membranes and also for many brackish water membranes, with
Eq. (5.49) and calculation of the logarithmic mean with Eq. (5.50), even for a
product recovery factor YE of up to 0.9, values for the average concentration factors
CFiEðintÞ and CFiEðlÞ are obtained that to a large part agree with each other,
corresponding to curve C in the graph, Fig. 5.19.
To calculate the average bulk concentration, it is also frequently proposed to use
the arithmetic mean of the feed concentration ci, F, E and the concentration of the
discharged concentrate ci, C, E.
The average bulk concentration as arithmetic mean is calculated from the com-
ponent concentrations and the element’s product recovery coefficient YE, while
considering the salt passage factor SPiE(t), as shown in Eq. (5.51).
 
1Y E SPiE ðtÞ
ci F,E þ ci,C,E ci F,E  1 þ 1Y E
ciFC,B,EðaÞ ¼ ¼ : ð5:51Þ
2 2
ci, C, E ¼ concentration of concentrate of component i [kg/m3] [g/l] [mg/l]
SPiE(t) ¼ salt passage factor of element for component i at temperature t []
RiE(t) ¼ salt rejection factor of element for component i at temperature t []
YE ¼ recovery coefficient of element []

Concentration factor
CF [ - ]
20
A
19
18
17
16
15 CF value at individual point x,y of
14 membrane element
13
12
11
Arithmetic average value of CF for B
10 membrane element
9
8
7 Integrated value and logarithmic
6 average value of CF for membrane
5 element
4
3 C
2
1
0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Recovery factor Y [ - ]

Fig. 5.19 Comparison of concentration factor CF values for membrane element calculation
384 5 Reverse Osmosis Membrane System: Core Process of SWRO

The mean arithmetic concentration factor CFiEðaÞ multiplied with the feed con-
centration ci, F, E to obtain the bulk concentration ciFC,B,EðtÞ then results under
consideration of the salt passage factor SPiE(t) and the salt rejection factor RiE(t) in
accordance respectively with Eqs. (5.51a) and (5.51b).

1Y E SPiEðtÞ
1þ 1Y E
CFiEðaÞ,SPi ¼ : ð5:51aÞ
2
1Y E ð1RiEðtÞ Þ
1þ 1Y E
CFiEðaÞ,Ri ¼ : ð5:51bÞ
2
CFiEðaÞ,SPi ¼ average concentration factor of element (arithmetic mean) consider-
ing salt passage []
CFiEðaÞ,Ri ¼ average concentration factor of element (arithmetic mean) consider-
ing salt rejection []
By neglecting the salt rejection of the separation membrane (RiE(t) ≌ 1),
Eq. (5.51c) is obtained.

1 þ 1Y
1
CFEðaÞ,Y E ¼ E
: ð5:51cÞ
2
CFEðaÞ,Y E ¼ average concentration factor of element (arithmetic mean) based on
recovery YE only []
Curve B in the graph of Fig. 5.19 shows how the concentration factor CFiEðaÞ,Ri
based on the arithmetic mean varies with the product recovery factor Y, and curve
C shows the same dependency for the integral and logarithmic means of the
concentration factors, CFiEðintÞ and CFiEðlÞ : respectively. Both curves are calculated
for a salt rejection factor, Ri, of 0.998. A comparison of the two curves shows that the
CF values calculated by forming the arithmetic mean show good agreement with the
values obtained by logarithmic and integral calculation up to a product recovery of
50%, but above this their divergence is substantial. For the arithmetic concentration
factor CFiEðaÞ,Ri , as product recovery rises increasingly higher values are noted, that
is the bulk concentrations at the membrane wall (and thus also the corresponding
osmotic pressures) calculated with Eqs. (5.51a)–(5.51c) are substantially higher than
is the case if the averages are determined in accordance with Eqs. (5.49) and (5.50).
Because for spiral-wound elements, the product recovery of the elements for
seawater is normally less than 10%, while for brackish water elements it is less than
15%, for calculating individual membrane elements and membrane modules the
arithmetic mean may find application. For CTA hollow-fibre elements, too, the
product recovery for each element for seawater desalination is usually 30–40%
and therefore below 50%. Likewise, for the seawater desalination stages of SWRO
plants, their product recovery normally does not exceed the range of 50–60% so the
simplified calculation using the arithmetic mean is possible. The situation is different
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 385

for the second pass of SWRO plants and other brackish water systems, which are
normally dimensioned for and operated with product recoveries of more than
80–90%. In these cases, as a minimum the logarithmic mean must be taken for the
calculation.
Like for calculating the concentration profiles for the salt content and for the salt
component i, also the osmotic pressure in the bulk solution and at the membrane wall
may be calculated using the average concentration factor and the concentration
polarization factor from the osmotic pressure of the feed to the membrane element
(Eqs. 5.52, 5.52a, and 5.52b). In this case, in accordance with Eq. (5.51c) and the
average concentration polarization factor on the basis of the salt content βsEðtÞ ,
CFEðaÞ,Y E may be taken as concentration factor (Eq. 5.48b).

π FC,M,EðtÞ ¼ π FC,B,EðtÞ  βsEðtÞ : ð5:52Þ

π FC,B,EðtÞ ¼ π FðtÞ  CFiEðaÞ,Y E : ð5:52aÞ

π FC,M,EðtÞ ¼ π FðtÞ  CFiEðaÞ,Y E  βsEðtÞ : ð5:52bÞ

π FC,M,EðtÞ ¼ average feed-to-concentrate osmotic pressure at membrane wall of


element at temperature t [bar]
πFC,B,EðtÞ ¼ average feed-to-concentrate bulk osmotic pressure of element at
temperature t [bar]

5.1.5.2.2 Concentration Polarization


For calculating membrane elements in practice, the concentration polarization in an
element may be characterized by calculating an average concentration polarization
factor βEðtÞ using a simplified form of Eq. (5.28). For this, the degree of concentration
polarization is determined from the ratio of the product flow FP,E(t) of the membrane
element to its average flow F WF,C together with a coefficient kM(t) specific to the
membrane (Eq. 5.53).
For calculating a membrane element or membrane module in practice, in the
following for the water and product flux as well as the production performance and
the product flow of these, instead of the symbols JW and FW, the symbols more
commonly used in practice, JP and FP, are used.
F P,E ðtÞ
k M ðtÞ 
βEðtÞ ¼ e F WF,C,E
: ð5:53Þ

kM(t) ¼ membrane coefficient at temperature t []


FP, E(t) ¼ product flow of membrane element at temperature t [m3/h]
F WF,C,E ¼ average feed-to-concentrate flow of element [m3/h]
If the average flow F WF,C,E through the element is calculated as the arithmetic
mean of the element’s feed flow FWF and discharge flow FWC in accordance with
Eq. (5.53a), then together with Eqs. (5.53), (5.54), and (5.54a), a relationship may be
386 5 Reverse Osmosis Membrane System: Core Process of SWRO

derived for which the average concentration polarization factor βEðtÞ depends solely
on the element’s product recovery, i.e. the product recovery factor YE (Eq. 5.55).

F WF,E  ð2  Y E Þ  
Y
F WF,C,E ¼ ¼ F WF,E  1  E : ð5:53aÞ
2 2
F P,E,ðtÞ
k M ðt Þ 
F WF,E ð2Y E Þ
βEðtÞ ¼ e 2 : ð5:54Þ

F P,E,ðtÞ
¼ YE: ð5:54aÞ
F WF,E
2Y E
βEðtÞ ¼ ekMðtÞ 2Y E ð5:55Þ

FWF, E ¼ feed flow of membrane element [m3/h]


YE ¼ recovery factor of element []
For 8" spiral-wound elements, depending on element type and configuration, like
spacer thickness and geometry, etc., the membrane coefficient kM(t) can take on
values from 0.70 to 0.80. With Eq. (5.53) in its simplified form and its derived form,
Eq. (5.55), however, it is only possible to approximate the complex relationships
arising for concentration polarization. This applies in particular for the feed-side
flow conditions and there, more specifically, regarding the influence of the salt
content and temperature of the feed flow to the element on the build-up and
composition of the solution boundary layer (see Sect. 5.1.5.1.3). For calculating
membrane elements in practice, however, the results of applying Eq. (5.55) are
sufficiently accurate to, for example, estimate the permissible product recovery of an
element if it is intended that the concentration polarization will not exceed a
specified maximum. The estimate for a permissible recovery factor YE at a specific
value of βEðtÞ is calculated with Eq. (5.56).

2  ln βEðtÞ
YE ¼ : ð5:56Þ
2  kM ðtÞ þ ln βEðtÞ

For the maximum value for the concentration polarization factor, βEðtÞ , of 1.2
stipulated by the membrane manufacturers, depending on the value selected for the
membrane coefficient kM(t), permissible product recoveries for spiral-wound
elements are calculated within a range of 18–21%, or a YE value of 0.18–0.21. For
seawater desalination in practice, though, elements in a membrane module are
usually dimensioned for substantially lesser product recoveries from around 8% to
10%. Consequently, also the concentration polarization is correspondingly lower
with βEðtÞ , taking on values from 1.07 to 1.10. For brackish water desalination, i.e. in
the second pass of a seawater desalination plant, the product recoveries of the
elements in a membrane module may attain efficiencies of between 10% and 15%.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 387

5.1.5.2.3 Element Capacity and Performance and Standard Test Conditions


Product (Solvent) Flow and Element Capacity
The production performance, i.e. water output, FP, E(t) of a membrane element at a
specified temperature t is calculated by Eq. (5.57) from its water permeability
referred to volume AE(t),V, the membrane area of the element SE, and the applied
net driving pressure NDPE(t). If the water permeability for a certain reference
temperature t0 is specified, additionally this value has to be multiplied by the
temperature correction factor fTCW to obtain the permeability and thus also produc-
tion performance at the desired temperature.

F P,E,ðtÞ ¼ J P,EðtÞ,V  SE ¼ AEðtÞ,V  SE  NDPEðtÞ


¼ AEðt0 Þ,V  f TCW  SE  NDPEðtÞ : ð5:57Þ

JP, E(t), V ¼ volumetric-based product flux of element at temperature t [m3/m2 h]


The product flux JP, E(t), V results from FP, E(t) or the water permeability coeffi-
cient AE(t), V and the membrane area SE of the element with Eqs. (5.57a) and (5.58).

F P,EðtÞ
J P,EðtÞ,V ¼ : ð5:57aÞ
SE
ΔpFC,EðtÞ  
J P,EðtÞ,V ¼ AEðtÞ,V  pF,E  pP,E   π FC,M,EðtÞ  π P,EðtÞ
2
¼ AEðtÞ,V  NDPEðtÞ : ð5:58Þ

JP, E(t), V ¼ product flux of element volumetric based at temperature t [m3/m2, h]


ΔpFC, E(t) ¼ total pressure differential of feed to concentrate of element at
temperature t [bar]
FP, E(t) ¼ product flow of membrane element at temperature t [m3/h]
AE(t), V ¼ volumetric water (solvent) permeability coefficient of membrane
element at temperature t [m3/m2hbar]
SE ¼ area of membrane element [m2]
π FC,M,EðtÞ ¼ average osmotic pressure of feed to brine at membrane wall at
temperature t [bar]
π P,EðtÞ ¼ average osmotic pressure of product at temperature t [bar]
For calculating a membrane element, the basic equation for the product flux has to
be extended by a correction factor to take account of its pressure loss ΔpFC, E. Also
for the element calculation, instead of the point-by-point differential osmotic pres-
sure, the averages of the osmotic pressures on the feed and product sides have to be
used. These are calculated from the osmotic pressure of the feed flow to the element
as shown in Eqs. (5.52), (5.52a), and (5.52b).
The pressure loss of the element ΔpFC, E(t) is made up of that for the new element
in its pristine state ΔpFC,Eclean ðtÞ and an additional differential pressure ΔpFC,Eoperation ðtÞ
that arises due to fouling and scaling of the element during operation. Both
388 5 Reverse Osmosis Membrane System: Core Process of SWRO

pressure losses together may not exceed the maximum permissible pressure loss
ΔpFC,Emax stipulated for the element (Table 5.2).

ΔpFC,EðtÞ ¼ ΔpFC,Eclean ðtÞ þ ΔpFC,Eoperation ðtÞ  ΔpFC,Emax : ð5:58aÞ

ΔpFC,Eclean ðtÞ ¼ pressure differential of feed to concentrate of clean element at


temperature t [bar]
ΔpFC,Eoperation ðtÞ ¼ pressure differential of feed to concentrate of operated element at
temperature t [bar]
ΔpFC,Emax ¼ max. allowable pressure differential of feed to concentrate of
membrane element
The average osmotic pressure at the membrane wall π FC,M,EðtÞ is calculated as
shown in Eq. (5.52b) from the osmotic pressure of the feed flow π F(t), the average
concentration factor of the element CFE , and the average concentration polarization
factor βEðtÞ. Under consideration of the pressure loss of the membrane element ΔpFC, E,
there then results the net driving pressure NDPE(t) as shown in Eq. (5.58b).

ΔpFC,EðtÞ  
NDPEðtÞ ¼ pF,E  pP,E   π F ðtÞ  CFE  βEðtÞ  π P,EðtÞ : ð5:58bÞ
2

The pressure loss ΔpFC,Eclean of a clean, scale-free membrane element may be


calculated either with the Darcy-Weisbach equation that finds application for pres-
sure loss calculations in fluid dynamics or with the help of empirical equations.
If the Darcy-Weisbach equation (Eq. 5.59) is used, though, also empirical criteria
specific for the flow pattern in the element are incorporated into the calculations.
Thus for spiral-wound elements, the frictional coefficient fΔp, the average effective
velocity υF,Ceff in the feed channel and its hydraulic diameter dH, and the specific
configuration of the spacer fabric inserted on the feed side in the membrane element
have to be taken into account. This fabric installed in the element’s feed channels
reduces its clearance cross-sectional area and volume.

f Δp  υ2F,C,eff  ρFðtÞ  l
ΔpFC,Eclean ¼
2  dH
 105 ðDarcy  Weisbach equationÞ: ð5:59Þ

ΔpFC,Eclean ¼ pressure loss of clean element [bar]


fΔp ¼ friction factor []
υF,C,eff ¼ effective feed to concentrate flow rate [m/s]
ρF(t) ¼ density of feed [kg/m3]
dH ¼ hydraulic diameter of flow channel [m]
l ¼ channel/element length [m]
By applying an empirical equation, the frictional coefficient fΔp for spiral-wound
elements is calculated from the Reynolds number Re and with values for the
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 389

coefficient a and the exponent b that are specific for the type and configuration of the
membrane element concerned (Eq. 5.59a) [48]. This equation applies for the
Reynolds number range from 100 to 1000. The Reynolds number Re itself is
obtained as shown in Eq. (5.59b).

f Δp ¼ a  Re b : ð5:59aÞ

ρFðtÞ  vF,C,eff  dH
Re ¼ : ð5:59bÞ
μF ðt Þ

a ¼ Reynolds number coefficient ¼ 6.23 for spiral-wound elements


b ¼ Reynolds number exponent ¼ 0.3 for spiral-wound elements
The average effective velocity υF,C,eff in the feed channel of a membrane element
is calculated as the quotient of the arithmetic average of the feed-to-concentrate flow
F WF,C in accordance with Eq. (5.53a) and the effective cross-sectional area of the
feed channel Aeff (Eq. (5.59c)).

F WF,C,E
υF,C,eff ¼ : ð5:59cÞ
Aeff
F WF,C,E ¼ average feed-to-concentrate flow of element [m3/h]
Aeff ¼ effective area of feed channel [m2]
Aeff is calculated by multiplying the width b and height hch of the feed channel
together with a factor to take account of its porosity ε (Eq. 5.59d). The porosity ε
results from the characteristics of the spacer fabric, namely the ratio of spacer
volume Vsp to the entire free volume of the feed channel VT (Eq. 5.59e). The volume
of the feed channel VT in turn results from multiplying the height hch, width b, and
length l of the channel (Eq. 5.59f).

Aeff ¼ b  hch  ε: ð5:59dÞ


V sp
ε¼1 : ð5:59eÞ
VT
V T ¼ hch  b  l: ð5:59fÞ

Aeff ¼ effective area of feed channel [m2]


hch ¼ channel height ¼ spacer thickness [m]
b ¼ channel width/length of membrane leaf [m]
ε ¼ porosity of feed channel []
Vsp ¼ spacer volume [m3]
VT ¼ free volume of channel [m3]
In order to obtain the hydraulic diameter dH of the feed channel of the membrane
element that is critical both for calculating the pressure loss ΔpFC,Eclean and for
obtaining the concentration polarization factor β(t) by means of the Sherwood
correlation (see Sect. 5.1.5.1.3), the specific parameters of the spacer fabric—spacer
390 5 Reverse Osmosis Membrane System: Core Process of SWRO

volume Vsp and spacer surface area Ssp—have to be known. With these two
parameters together with the free volume of the feed channel VT as well as its
surface area Ssp, by substitution into Eq. (5.59h) there, then results, with
Eq. (5.59g), the hydraulic diameter dH [48].

Sfc ¼ 2  ðhch þ bÞ  l: ð5:59gÞ


 
4  V T  V sp
dH ¼ : ð5:59hÞ
Sfc  Ssp

Sfc ¼ surface of channel


Ssp ¼ surface of spacer [m2]
From the various geometric parameters of the feed channel and the spacer fabric
as described above, Eq. (5.59i) can then be derived for the hydraulic diameter dH,
which is only still dependent on the feed channel’s porosity ε, its height hch, and the
specific surface area of the spacers Sv,sp. Sv,sp is then obtained from the ratio of the
spacer parameters of surface area Ssp to spacer volume Vsp, or as shown in
Eq. (5.59j), from one or other of these parameters if the spacer’s filament thickness
dF is known.

4ε
dH ¼ : ð5:59iÞ
2
hch þ SV,sp  ð1  εÞ

Ssp 4
SV,sp ¼ : ð5:59jÞ
V sp dF

Sv,sp ¼ specific surface of spacer [m2]


dF ¼ filament thickness of spacer [m]
For calculating the pressure loss of spiral-wound membrane elements in practice,
simpler empirical relationships are used. For these, the differing designs of the
elements regarding spacer thickness and the type of spacer fabric are considered
by applying element-specific coefficients and exponents (Eqs. 5.60 and 5.60a).
However, when these equations are used without correction functions for tempera-
ture and salt content for the specific coefficients and exponents, these influences, that
are considered in the Darcy-Weisbach algorithms, are neglected.12
 a
ΔpFC,Eclean ¼ 6:895  104  4:4  F WF,C,E : ð5:60Þ
h  ia
Y
ΔpFC,Eclean ¼ 6:895  104 4:4  F WF,E  1  E : ð5:60aÞ
2
for 8-inch elements:

12
Source: DuPont Filmtec Reverse Osmosis Membranes Technical Manual Table 28.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 391

a ¼ element-specific exponent ¼ 1.3–1.7


ΔpFC,Eclean ¼ pressure loss of clean element [bar]
F WF,C,E ¼ average feed-to-concentrate flow of element [m3/h]
FWF, E ¼ feed flow of membrane element [m3/h]

Salt Flux, Salt Flow, and Membrane Rejection


Likewise for calculating the salt flux JSi, E(t) and the salt flow FSi, E(t) of a component
i, the equation for point-by-point modelling for salt transport through a separation
membrane (Eq. 5.42) has to be adjusted to match the conditions in a membrane
element by applying mean values for the bulk concentration and the concentration of
component i at the membrane wall instead of point-by-point concentrations
(Eqs. 5.61, 5.61a, 5.61b, and 5.61c)
 
J Si,EðtÞ ¼ BiEðtÞ  ciFC,M,EðtÞ  ciP,EðtÞ : ð5:61Þ
 
J Si,EðtÞ ¼ BiEðtÞ  ci F,E  CFiE  βiEðtÞ  ciP,EðtÞ : ð5:61aÞ
 
J Si,EðtÞ ¼ BiEðt0 Þ  f TCS  ci F,E  CFiE  βiEðtÞ  ciP,EðtÞ : ð5:61bÞ
 
F Si,E,ðtÞ ¼ BiEðtÞ  SE  ci F,E  CFiE  βiEðtÞ  ciP,EðtÞ : ð5:61cÞ

ciP,EðtÞ ¼ average product concentration of component i at temperature t [kg/m3]


[g/l] [mg/l]
JSi, E(t) ¼ salt flux of element for component i at temperature t [kg/m2, h]
FSi, E(t) ¼ salt (solute) flow for element of component i at temperature t [kg/h]
BiE(t) ¼ salt permeability coefficient of element for component i at temperature
t [m/h]
BiEðt0 Þ ¼ salt permeability coefficient of element for component i at reference
temperature t0 [m/h]
The equations for salt rejection Ri,E(t) (Eq. 5.62) and salt passage SPi,E(t) (Eq. 5.63)
for a membrane element have the same basic form as is the case for point-by-point
membrane modelling (Eqs. 5.23, 5.24, and 5.27). However, the value for net driving
pressure NDPE(t) is now calculated in accordance with Eq. (5.58b) with averaging of
the osmotic pressure at the membrane wall and in the permeate. Also to be consid-
ered is the membrane element’s differential pressure ΔpFC, E.
 1  1
BiEðtÞ 1
RiEðtÞ ¼ 1þ ¼ 1þ : ð5:62Þ
AEðtÞ,V  NDPEðtÞ PiEðtÞ  NDPEðtÞ
 1
AEðtÞ,V  NDPEðtÞ  1
SPiE ðtÞ ¼ 1þ ¼ 1 þ PiEðtÞ  NDPEðtÞ : ð5:63Þ
BiEðtÞ
392 5 Reverse Osmosis Membrane System: Core Process of SWRO

ciP,EðtÞ ¼ ci,F,E  CFiE  βiEðtÞ  SPiE ðtÞ


 
¼ ci F,E  CFiE  βiEðtÞ  1  RiEðtÞ : ð5:64Þ

The mean concentration of component i ciP,EðtÞ in the product water of a


membrane element results from the feed concentration ciF,E to the element and
knowledge of the mean concentration factor CFiE , the mean concentration polariza-
tion factor βi,EðtÞ, and the factor for salt passage SPiE(t) or salt rejection RiE(t) as shown
in Eq. (5.64).

Determination of Permeation Coefficients AE,V and BiE and Salt Rejection RiE
The production performance of a membrane element FP,E(t) is governed by its water
permeability coefficient AE(t),V plus the feed conditions at which it is operated, like
pressure pF,E, salt concentration or concentration ci, F of the component i, and
temperature tF as well as, on the product side, product recovery YE and pressure
pP,E. The permeability coefficient AE(t),V of a membrane element is determined
empirically by subjecting it to defined values of the above mentioned feed and
operating conditions using a solution of defined concentration and composition.
From the resulting measurement of the element’s production performance FP,E(t), the
water permeability coefficient can be calculated with Eq. (5.65).

F P,EðtÞ
AEðtÞ,V ¼ h  i : ð5:65Þ
ΔpFC,EðtÞ
SE  pF,E  pP,E  2  π FðtÞ  CFE  βEðtÞ  π P,EðtÞ

The osmotic pressures π F(t) of the feed solution and the product water π P,EðtÞ are
calculated from the component concentrations as described in Chap. 3, Sect. 3.2.3.3,
Eqs. (3.115)–(3.115g), together with Eqs. (3.116a)–(3.118) or also according to
Eqs. (5.40) and (5.40a).
Under the same test conditions for the membrane element, also the salt perme-
ability coefficient BiE(t) of a component i of the solution to be treated can be
determined (Eqs. 5.65a and 5.65b). For this, as well as the element’s production
performance FP,E(t), the concentration ciP,EðtÞ of the component i in the product water
has to be measured. Additionally, either the mean bulk concentration in the element
ci FC,B,EðtÞ together with the mean concentration polarization factor βEðtÞ or the feed
concentration ciF,E and the product CFE  βEðtÞ have to be known.

F P,EðtÞ,V
BiEðtÞ ¼ c : ð5:65aÞ
i FC,B,E ðt ÞβE ðtÞ
SE  ciP,EðtÞ 1
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 393

F P,EðtÞ,V
BiEðtÞ ¼  : ð5:65bÞ
ci F,E CFE βE ðt Þ
SE  ciP,EðtÞ 1

For calculating AE(t),V and BiE(t), depending on how the mean has been calculated,
the mean concentration factor CFE is obtained from Eqs. (5.49), (5.50), (5.51a),
(5.51b), or (5.51c). The concentration polarization factor βEðtÞ is calculated from the
recovery factor YE using Eq. (5.55).
Determination of the intrinsic salt rejection RiE(t) of the membrane element is
done by measuring or calculating the same concentration parameters as used for
determining BiE(t).

ciP,EðtÞ
RiEðtÞ ¼ 1  : ð5:66Þ
ci FC,B,EðtÞβEðtÞ

ciP,EðtÞ
RiEðtÞ ¼ 1  : ð5:66aÞ
ci F,E  CFE βEðtÞ

When conducting the tests to determine the performance parameters of a mem-


brane element, it must be borne in mind that new membranes subjected to pressure
undergo physical changes in the membrane separation layer due to the membrane
being compacted. The membrane’s water permeability decreases while its salt
rejection increases. This phase up to stabilization of the membrane behaviour until
a constant level is attained may take less than an hour or up to several hours
depending on membrane type and how it has been manufactured. For this reason,
in order to obtain reliable results, a membrane test should be continued until the
values measured for performance and salt rejection have stabilized and no longer
change during the test period. The measuring results for a membrane element’s
production performance and salt rejection so obtained are termed stabilized values.

Ageing of Membranes and Long-Term Performance


All of the calculation equations described above for the water and salt transport at
membranes and within membrane elements apply for ideal conditions, i.e. for new or
as-new membranes that have not undergone physical or chemical structural changes.
During actual operation of a desalination plant, though, its membrane elements are
subject to ageing and this modifies the mass transport characteristics and salt
rejection of the separation membranes. Factors that influence the membranes’ ageing
process and modify their performance characteristics are:

• membrane compaction and how it is affected by operating pressure and tempera-


ture during desalination
• natural ageing of the plastic materials, in particular the membrane separation layer
• structural chemical changes of the membrane separation layer, for example, due
to oxidation and hydrolysis
394 5 Reverse Osmosis Membrane System: Core Process of SWRO

• irreversible coating of the membrane surface by inorganic or organic substances


that cannot be removed by chemical membrane cleaning.

However, critical for the degree of age-related changes of the membrane


properties is not only the nature of the factors that promote ageing, but also the
duration of their respective actions on the membranes.
When dimensioning membrane desalination systems in practice, the changes of
the membrane element’s water and salt permeabilities have to be taken into account
throughout their operating and service lifetime. This is done by applying correction
factors for membrane permeability, these being the correction factor f FP,E,τMOp for the
change in water permeability AE(t),V (Eqs. 5.67 and 5.67a), and the correction factor
f FSi,E,τMOp for the salt permeability BiE(t) (Eqs. 5.68 and 5.68). Both these factors are
empirical parameters that depend on the specific operating conditions of the mem-
brane elements, like pressure, temperature, chemical influences, and quality of the
feed as well as membrane operating parameters like membrane flux and product
recovery. The initial membrane permeability or initial product flux and salt flux of a
membrane element are multiplied by these correction factors so as to obtain their
values after a defined membrane element operating time.
These correction factors are determined by the membrane manufacturers that are
derived primarily from empirical values for operational desalination systems, that is
by evaluating their operating data but also on the basis of long-term tests of
membrane modules.
The water permeability of membrane elements decreases as they age, which
means that the correction factor f FP,E,τMOp for the initial water permeability coefficient
AE(t), V is less than 1 and continues to drop as the membrane’s age increases.

AEðtÞ,V,τMOp
f FP,E,τMOp ¼ ¼< 1: ð5:67Þ
AEðtÞ,V

AEðtÞ,V,τMOp ¼ AEðtÞ,V  f FP,E,τMOp : ð5:67aÞ

AEðtÞ,V,τMOp ¼ water permeability coefficient of volumetric-based membrane


element at temperature t and operation time of element τOp [m3/m2hbar]
AE(t), V ¼ initial water permeability coefficient of volumetric-based membrane
element at temperature t [m3/m2hbar]
f FP,E,τMOp ¼ product permeability coefficient correction factor (flow factor) of
element at membrane operating time τOp []
The product flux J P,EðtÞ,V,τMOp and production performance F P,E,ðtÞ,τOp of a mem-
brane element after an operating time τMOp, E and taking into account the correction
factor for product permeability f FP,E,τMOp may then be calculated with Eqs. (5.67b)
and (5.67c).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 395

J P,EðtÞ,V,τMOp ¼ AEðtÞ,V,τOp  NDPEðtÞ ¼ AEðtÞ,V  f FP,E,τMOp  NDPEðtÞ : ð5:67bÞ

F P,E,ðtÞ,τMOp ¼ J P,EðtÞ,V,τMOp  SE ¼ F P,EðtÞ  f FP,E,τMOp : ð5:67cÞ

J P,EðtÞ,V,τMOp ¼ volumetric product flux of element at temperature t at an operation


time τOp [m3/m2, h]
F P,EðtÞ,τMOp ¼ product flow of element at temperature t for an operation time τOp
3
[m /h]
FP, E(t) ¼ initial product flow of element at temperature t [m3/h]
Salt permeability BiEðtÞ,τi and salt passage SPiE(t) increase with increasing mem-
brane age τMOp, E, which means that the correction factor f FSi,E,τMOp for the initial salt
permeability BiE(t) is greater than 1 and increases with the operating time of a
membrane element (Eqs. 5.68 and 5.68a).

BiEðtÞ,τOp
f FSi,E,τMOp ¼ > 1: ð5:68Þ
BiEðtÞ

BiEðtÞ,τOp ¼ BiEðtÞ  f FSi,E,τMOp : ð5:68aÞ

BiEðtÞ,τMOp ¼ salt permeability coefficient of element for component i at tempera-


ture t and operation time of element τMOp [m/h]
BiE(t) ¼ initial salt permeability coefficient of membrane element at temperature
t [m/h]
f FSi,E,τMOp ¼ salt permeability coefficient correction factor of element for compo-
nent i at an element operation time τMOp []
The salt flux of a membrane element as a function of its age τMOp, E results from
Eqs. (5.68b)–(5.68d).

J Si,E,ðtÞ,τMOp ¼ BiEðtÞ,τMOp  Δci,EðtÞ ¼ BiEðtÞ  f FSi,E,τMOp  Δci,EðtÞ : ð5:68bÞ

Δci,EðtÞ ¼ ci F,E  CFiE  βiEðtÞ  ciP,EðtÞ : ð5:68cÞ

J Si,E,ðtÞ,τMOp ¼ J Si,EðtÞ  f FSi,E,τMOp : ð5:68dÞ

Δci, E(t) ¼ feed-to-product concentration gradient of component i in element


[kg/m3] [g/l]
JSi,EðtÞ,τOp ¼ salt flux of element of component i at temperature t and element
operation time τMOp [kg/m2, h]
The product permeability correction factor f FP,E,τMOp is also termed by the
membrane manufacturers as the membrane flux retention coefficient (MFRC).
Essentially, this characterizes the change of water permeability of a membrane
element as a result of membrane compaction and natural ageing. How this changes
with membrane age τOp can be approximately depicted by an algorithm in accor-
dance with Eqs. (5.69) and (5.69a). The exponent or coefficient m characterizes the
396 5 Reverse Osmosis Membrane System: Core Process of SWRO

curve’s slope that expresses the dependency of the MFRC on the membrane’s
operation time τMOp, E. The membrane compaction coefficient m depends on the
membrane type as well as the pressure and temperature to which it is subjected.

F P,E,τMOp AEðtÞ,V,τMOp
¼ ¼ τMOp,E m ¼ MFRC: ð5:69Þ
F P,EðtÞ AEðtÞ,V

log AEðtÞ,V,τMOp ¼ log AEðtÞ,V  m  log τMOp,E : ð5:69aÞ

τMOp, E ¼ operation time (lifetime) of membrane element [h]


m ¼ membrane compaction coefficient []
MFRC ¼ membrane flux retention coefficient []
Following the first application of pressure to new membrane elements, an
approximately constant compaction coefficient m will only become established
after a certain running in or stabilization period. Depending on the membrane type
and how it is manufactured, this may take the first few hours or even days after the
commencement of pressure application. During this period, the m value is signifi-
cantly lower than it will be for the ensuing membrane lifetime, so that then the
membrane’s water permeability degrades considerably more rapidly than is the case
during subsequent operation (Fig. 5.20). This stabilization period is accompanied by
an increased rate of compaction and the membrane element’s salt rejection perfor-
mance may improve. Depending on the membrane type and its manufacturing
process, for individual membrane elements this stabilization period may last a few
hours, but for larger desalination plants with many elements it may extend to as

Water permeability loss


by compaction
AτOp /Aτi
1.1

CTA membrane at feed pressure of 60 bar


1.0

Temperature [°C]

10 °C
0.9 15 °C

20 °C

25 °C

0.8
30 °C

0.7
0 1 2 3
Membrane operation time τMOp [years]

Fig. 5.20 Membrane compaction and aging: water permeability loss of CTA membrane as a
function of temperature
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 397

much as a week. This is particularly the case for seawater desalination membranes
due to their operation at substantially higher pressures compared with brackish water
desalination.
The dependency of the age-related degradation of water permeability of a CTA
seawater desalination membrane on temperature and pressure is shown in Figs. 5.20
and 5.21.
The operating temperature of a membrane desalination plant greatly influences
the drop in the membrane’s water permeability and thus the plant’s production
performance as the age of its membrane elements increases. Membrane compaction
is irreversible, which means that a reduction in performance caused by lengthy
operation at a higher feed temperature will remain even during subsequent operation
at a lower temperature. When dimensioning a desalination plant whose operating
temperature fluctuates over wide ranges with peak temperatures of 30  C and more,
this effect has to be taken into account accordingly.
The variation of the salt permeability of a cellulose triacetate (CTA) membrane
over its operational lifetime τMOp and thus also its potential membrane lifetime is
determined in particular by chemical influences. These are the oxidixing effects of
the chlorine dosed for disinfecting the feed coupled with the hydrolysing action of
the feed pH (see Sects. 5.1.3.2 and 5.1.3.2.2). The reduction of the salt permeability
of a CTA hollow-fibre membrane and therefore the corresponding correction factor
f FSi,E,τMOp for the initial salt permeability coefficient BiEðtÞ,τi as a function of its
contact time tc with an oxidizing and hydrolysing impact may be calculated using
Eq. (5.70) together with Eqs. (5.70a)–(5.70e).

Water permeability loss


by compaction
AτOp /Aτi
1.0

CTA membrane at temperature of 25°C

0.9

Pressure [bar]

50
55
60
65

0.8
0 1 2 3
Membrane operation time τMOp [years]

Fig. 5.21 Membrane compaction and ageing: water permeability loss of CTA membrane as a
function of pressure
398 5 Reverse Osmosis Membrane System: Core Process of SWRO

Using Eq. (5.70a), for a given maximum increase of the salt permeability of the
CTA membrane, it is possible to determine the permissible contact time tc [36].

Bi,τOp ,
log ¼ K  τc : ð5:70Þ
Bi, , τi
Bi,τOp ,
log Bi, , τi
τc ¼ : ð5:70aÞ
K
K ¼ K H þ KO: ð5:70bÞ
6, 400
log K H ¼ þ 0:80  pH þ 11:0: ð5:70cÞ
T
1, 850
log K O ¼ 1:072  log cCl2  þ 0:89  pH  4:246: ð5:70dÞ
T
14,740
þ1:842pH
K ¼ 1  1011  e T þ 5:675  105  cCl2 1:072
4,260
e T þ2:049pH ð5:70eÞ

K ¼ B value increase rate coefficient [h1]


KH ¼ B value increase rate coefficient by hydrolysis [h1]
KO ¼ B value increase rate coefficient by oxidation [h1]
cCl2 ¼ chlorine contact concentration [mg/l Cl2]
τc ¼ time of chlorine contact with membrane [h]
T ¼ temperature [K]
If for a chlorine contact concentration cCl2 of 0.5 mg/l and a pH of 6.5, the value of
the salt permeability increases to not more than double, i.e. if an initial salt rejection
rate of 99.8% reduces to not more than 99.6%, then under these conditions a contact
time tc of 30,000 h, or 3.4 years, is permissible for the CTA membrane. Under
continuous contact, this also corresponds to the maximum lifetime τMl max of the
membrane. If this contact is not continuous, for example with intermittent chlorine
dosing, the increase of the membrane’s lifetime is equal to the period for which it is
not in contact with the oxidizing medium (for more details, see Sect. 5.1.3.2,
Eqs. 5.4, 5.5, 5.7, and 5.8). However, the permissible contact time tc may substan-
tially decrease if there is a catalytic reaction in the presence of heavy metals [39] (see
Sect. 5.1.3.2.2).
For a chlorine contact concentration of 0.5 mg/l and a temperature of 20  C, the
dependency of the correction factor f FSi,E,τMOp , i.e. the ratio of the initial to the
BiEðtÞ,τMOp
operational salt permeability BiEðtÞ , on the maximum contact time tc of a CTA
membrane is plotted for various feed pH values in Fig. 5.22. This graph clearly
shows the strong influence of hydrolysis on the lifetime of CTA membranes.
The sensitivity to the oxidizing influences of chlorine and chlorinated
components is much greater for polyamide TFC membranes than is the case for
cellulose acetate membranes. Even trace concentrations of chlorine may result in
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 399

Increase of Salt
permeability coefficient
BτOp /Bτi
10,0
pH = 7.5 pH = 7.0

9,0 Chlorine contact concentration = 0.5 mg Cl2/l


Temperature = 20 °C pH = 6.5

8,0

7,0

6,0

5,0

4,0

3,0
pH = 6.0
2,0

1,0
0 10.000 20.000 30.000 40.000 50.000 60.000 70.000 80.000 90.000 100.000
Max. Contact time τc [h]

Fig. 5.22 CTA membrane: salt permeability increase at 0.5 ppm chlorine contact concentration for
various values of pH

substantial changes in the chemical structure and thus bring about a bigger impact, in
particular of the salt permeability of these membrane types (see Sect. 5.1.3.2.1).
With Eqs. (5.71) and (5.71a), estimates can be made of how the quotient of the
BiEðtÞ,τMOp
operational and initial salt permeability BiEðtÞ changes for specific chlorine contact
concentrations and what maximum permissible contact time τc for the polyamide
membrane then results [51]. In this case also catalytic effects could substantially
shorten the contact time.
 
Bi,τOp , 1:6103 c0:5
Cl τ c
¼e 2
: ð5:71Þ
Bi, , τi
Bi,τOp ,
ln Bi, , τi
τc ¼ 3 pffiffiffiffiffiffiffi : ð5:71aÞ
1:6  10  cCl2

Figure 5.23 shows how the salt permeability of a polyamide TFC membrane
when plotted against maximum contact time increases with chlorine contact
concentrations from 0.05 to 0.2 mg/l Cl2.
For a chlorine concentration as low as 0.05 mg/l, after just 2000 h, or 0.23 years,
of continuous contact, the salt permeability, i.e. the passage of salt through the
membrane, is doubled and for a contact concentration of 0.2 mg/l Cl2 it is four times
as high. For this reason, even with intermittent dosing of chlorine and chlorinated
400 5 Reverse Osmosis Membrane System: Core Process of SWRO

Increase of Salt
permeability coefficient
BτOp /Bτi
5

Cl2 - Concentration
[mg/l]

0.2
4
Polyamide- TFC - Membrane of UTC-70 type

0.1

2 0.05

1
0 200 400 600 800 1.000 1.200 1.400 1.600 1.800 2.000 2.200 2.400
Max. Contact time τC [h]

Fig. 5.23 Polyamide TFC membrane: increase in salt permeability due to contact with chlorine

substances, it would hardly be possible to obtain a membrane service life of 5 years


that would be necessary for economic operation of an industrial-scale membrane
desalination plant. It is therefore imperative that the contents of chlorine and
oxidizing chlorinated compounds as well as other oxidizing components be reduced
in the feed to the membrane elements to the limit of detectability, as stipulated by the
membrane manufacturers.
For polyamide spiral-wound elements, the loss of water permeability with age as
expressed by the permeability correction coefficient f FP,E,τMOp is calculated by the
manufacturers of these membranes using Eq. (5.72). The parameter fΔA is a measure
of the annual drop of the water permeability and is therefore termed the permeability
decline factor. The quantity 1  fΔA is also designated by various membrane
manufacturers as the fouling factor or flow factor ff.

f FP,E,τMOp ¼ ð1  f ΔA ÞτMOp,E ¼ f f τMOp,E : ð5:72Þ

fΔA ¼ product permeability decline factor of element []


ff ¼ flow factor []
The change of salt permeability with membrane age, which in this case is the
increase of the salt permeability correction coefficient f FSi,E,τMOp , is calculated from
the factor for the annual rise fΔB in accordance with Eq. (5.73).

f FSi,E,τMOp ¼ 1 þ f ΔB  τMOp,E : ð5:73Þ

fΔB ¼ salt permeability increase factor of element []


5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 401

When operating a membrane desalination plant, however, its performance doesn’t


change only due to ageing of the membrane elements, termed irreversible fouling,
but also due to coating of the membrane surfaces with inorganic and organic
particulate pollutants and through biological growth in the elements. To a great
extent, this fouling can be removed by flushing and chemical treatment of the
membranes and is thus termed reversible fouling. The extent to which the
membranes’ performance deteriorates due to reversible fouling in the period
between membrane cleaning runs depends on the quality of the feed to the mem-
brane desalination train, i.e. the nature of the upstream cleaning process and its
effectiveness as well as the frequency with which such a cleaning run is carried out
and its efficiency (see Sect. 5.5.2.5).
The factors quoted by the membrane manufacturers for the change in permeabil-
ity fΔA and fΔB, though, apply only for the irreversible fouling of the membranes due
to their ageing and then also only for clean and treated membranes. Consequently,
they are also designated f ΔAclean and f ΔBclean .
The membrane manufacturers specify specific bandwidths for the permeability
correction coefficients f ΔAclean and f ΔBclean for “clean” membranes, with these
stipulated values being oriented to, among others, how the raw water is extracted
and its further pretreatment, that is to the anticipated quality of the feed to the RO
train and possible fluctuations of this. This also takes into account the aspect of
irreversible coating of the membranes or also increased chemical influencing of the
membrane structure in the event of more frequent and more intensive chemical
cleaning for problematical raw water conditions.
In this context, “clean” membrane is to be understood as the performance of a
membrane element directly after it has been subjected to an efficient membrane
cleaning process.
Compiled in Table 5.3 are values of the factors stipulated by the membrane
manufacturers for polyamide spiral-wound elements for the annual change of
water permeability f ΔAclean and salt permeability f ΔBclean .

Table 5.3 Product permeability decline and salt permeability increase factors of clean polyamide
TFC membrane elements
f ΔAclean ff f ΔBclean
Source of RO feed/pre-treatment %/
type year year1 year1 %/year year1
Open intake/conventional 7–12 0.07–0.12 0.88–0.93 7–10 0.07–0.10
Open intake/UF or MF 5–10 0.05–0.10 0.90–0.95 7–10 0.07–0.10
Sea well/sand filtration 5–10 0.05–0.10 0.90–0.95 7–10 0.07–0.10
1. Pass RO product 3–7 0.03–0.07 0.93–0.97 3–5– 0.03–0.05–
10a 0.1a
a
At high pH
402 5 Reverse Osmosis Membrane System: Core Process of SWRO

The bandwidths of these values have been compiled from the data of various
manufacturers. It is quite possible that, for certain operating conditions, specific
membrane manufacturers could stipulate lower or higher value ranges.
The factors fΔA and fΔB depend, like the m value in Eq. (5.69), on temperature and
pressure as well as the other operating conditions of a membrane element. The data
in Table 5.3 are therefore to be understood as input values that are matched to these
conditions in the membrane manufacturers’ design software in accordance with the
specific design conditions there specified together with the manufacturers’ experi-
ence of these conditions. The ability to forecast the long-term behaviour of his
membranes under varying operating conditions is a key component of a membrane
manufacturer’s know-how. For example, when dimensioning his membranes, one
manufacturer of TFC spiral-wound elements—DOW/DuPont Filmtec—only
corrects the product permeability and states that the salt passage of his membranes
does not change as they age.
The relationship given by Eq. (5.74) is derived from Eq. (5.72). If the maximum
permissible loss of water permeability or production performance of an element,
i.e. the maximum value of the permeability correction coefficient f FP,E,τMOp , is
stipulated, this can be used to calculate the associated lifetime of the membrane
element τMOp, E.

log f FP,E,τMOp
τMOp,E ¼ : ð5:74Þ
log ð1  f ΔA Þ

Also from Eq. (5.73) for the increase of salt passage as the membrane ages, by
specifying its permissible rise, it is possible to estimate the maximum age the
membrane element can then attain (Eq. 5.75).

f FSi,E,τMOp  1
τMOp,E ¼ : ð5:75Þ
f ΔB

The results of such calculations for various values of fΔA and fΔB are graphed in
Fig. 5.24. The data shown in the graph are based on the assumption that the
membrane’s loss of performance due to fouling is reversible and can be almost
completely eliminated by chemical treatment or, alternatively, that the irreversible
coating of the membranes has already been taken into account by the factors f ΔAclean
and f ΔBclean quoted by the manufacturers.
If for an SWRO plant with open intake extraction and conventional pretreatment
with a correction factor for water permeability, fΔA, of 0.1 and one for salt perme-
ability, fΔB, of 0.07, the production performance does not drop by more than 30%
and the product salt content does not rise by more than 20%, there results for a TFC
spiral-wound element a presumed membrane service life of approximately 3 years
under the aspects of salt permeability rise and loss of production performance.
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 403

Fig. 5.24 Polyamide TFC membrane: salt and water permeability correction factor vs. membrane
lifetime

However, during operation of a membrane desalination plant in practice, nor-


mally the reduction of the membrane’s water permeability is compensated by a
corresponding increase in operating pressure, i.e. of the net driving pressure, NDP.
In this way, as shown by Eqs. (5.62) and (5.63), also the membranes’ salt perme-
ability is affected in that it is reduced while its salt rejection increases. But,
additionally, when actually operating a membrane desalination plant, in practice
the membrane elements with which it is initially equipped do not remain installed
until a correspondingly reduced performance is attained. In accordance with a
membrane replacement schedule that is specified when the plant is designed, an
average membrane age is set that, depending on the stipulated rate of membrane
replacement, is usually attained after an operating time of around 5–10 years, and
this is then kept constant by appropriately adjusting the membrane change-out cycle.
For this reason, a differentiation is made between the above stated membrane age
τMOp,Ei of individual elements and an average membrane lifetime (AMLT) τMOp,Ø,S
of the membrane system. The latter is calculated as the quotient of the sum of the
membrane ages of all membrane elements installed in the system and the number of
these elements NS (Eq. 5.76) (see Sect. 5.2.3.3.1).
404 5 Reverse Osmosis Membrane System: Core Process of SWRO

i¼N ES
P
τMOp,Ei
τMOp,S i¼1
τMOp,∅,S¼AMLT¼ ¼ : ð5:76Þ
N ES N ES

τMOp, ∅ , S¼ AMLT ¼ average membrane lifetime of system at τSOP, AMLT


[years]
τMOp, S ¼ membrane age of system at τSOP, AMLT [years]
τMOp,Ei ¼ membrane age of individual element i at τSOP, AMLT [years]
N ES ¼ number of membrane elements in the membrane system []
τSOP, AMLT ¼ RO system operation time with the AMLT calculation [years]

Standard Test Conditions of Commercial Membrane Elements


The production performance and salt rejection of commercial RO membrane
elements are specified by their manufacturers on the basis of “standard test
conditions”. The type and scope of the parameters for these conditions correspond
in principle to the same operating conditions under which tests are conducted of
membrane elements to determine their permeability coefficients and salt rejection
rates, these being namely defined values for:

• feed pressure
• product backpressure
• temperature
• product recovery
• concentration of feed solution and types of ionogenic salt components.

Compiled in Table 5.4 are the standard test conditions with which the
manufacturers of polyamide spiral-wound element and hollow-fibre CTA membrane
elements specify the performance characteristics of their seawater and brackish
water membrane elements.
Table 5.4 also shows the classification of spiral-wound elements into various
categories depending on their production performance and salt rejection rates. The
classification shown is used in a comparable way by all leading manufacturers of this
type of membrane elements.
These standard tests are conducted under atmospheric pressure on the product
side and thus without considering the product pressure. The membrane elements are
supplied with a standard test solution which, for characterizing the total salt rejection
of the membranes, contains sodium chloride, NaCl, as salt component at a defined
concentration. If the rejection of a certain component i, for example boron, B, is to be
characterized, this will likewise be referred to a specified concentration of this
component in the feed solution. Additionally, specific feed conditions may be
defined, such as a certain pH in the case of boron.
For seawater elements, the manufacturers of spiral-wound modules define largely
identical standard test conditions irrespective of the category to which the element is
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 405

Table 5.4 Standard test conditions for commercial seawater and brackish water RO membrane
elements
Standard test conditions
Tempe- Element
Salt content Pressure rature recovery
cSt,NaCl,F,E pSt,F,E tSt,F,E YSt,E
Membrane element type and mg/l mg/l B at

category NaCl pH Bar C %
Seawater RO elements
Spiral-wound composite 32,000 5 at 8 55.2 25 8–10
elements
• Normal
• High rejection
• High flux/low energy
• Low fouling
Hollow-fiber CTA 35,000 – 53.9 30
elements
Brackish water elements
Spiral-wound composite 1500– 5 at 8–10 15.5 25 15
elements 2000
• Normal
• High rejection
• Higher flux/lower 10.3
rejection

assigned. A few manufacturers, though, refer the standard value for an element’s
product recovery to 10% in contrast to the 8% value of most of the other providers.
As shown by Table 5.4, the standard test conditions for hollow-fibre CTA
membrane elements differ appreciably from the test conditions for spiral-wound
elements with regard to the NaCl concentration of the test solution as well as the
applied pressures and the elements’ product recovery rates.
For the brackish water spiral-wound elements, all manufacturers base their
standard test conditions on the same product recovery rate of 15%. Manufacturers
differ in the NaCl concentration of the test solution and also the applied pressure.
Depending on the type of element, however, test pressures vary, but are not the same
for comparable membrane types of the manufacturers.
The standard temperature of 25  C is the same for all membrane manufacturers.
Compiled for commercial RO membrane elements for seawater and brackish
water desalination in Tables 5.37, 5.38, and 5.39 in Annexes 5.A2, 5.A2.1, and 5.A3
and broken down by membrane type—polyamide spiral-wound composite, polyam-
ide spiral-wound nanocomposite, and CTA hollow fibre—as well as the respective
membrane categories are the standard test conditions and performance parameters as
quoted by individual manufacturers for their membrane elements.
From the data there specified for the production performance FPE(St) and the salt
rejection RENaCl(St) of a membrane element with a membrane area SE, knowledge of
406 5 Reverse Osmosis Membrane System: Core Process of SWRO

the associated standard conditions for pressure pFE(St), temperature tFE(St), NaCl
concentration cNaCl(St),FE and the element’s product recovery YE(St), and with the
aid of Eq. (5.65) for its water permeability coefficient AE(St),V together with
Eqs. (5.65a) and (5.65b), it is also possible to determine the salt permeability
coefficient BNaCl(St),E (see Sect. 5.1.5.2.3).
For the various commercial seawater and brackish water membranes with their
performance figures under standard conditions as listed in Tables 5.37, 5.38, and
5.39 in Annexes 5.A2, 5.A2.1, and 5.A3, the coefficients for their water and salt
permeabilities AE(St),V and BNaCl, E(St) derived from these parameters are shown in
Table 5.5.
When specifying the salt rejection of their membrane elements under standard test
conditions, manufacturers quote differing figures for the salt rejection of the
membranes. Distinctions are made between “minimum salt rejection”, “nominal
salt rejection”, and “stabilized salt rejection”. Also the specifications differ regarding
the duration of the test during which they are determined. Calculation of the
permeability coefficients shown in Table 5.5 is based either on the “nominal salt
rejection” quoted by the membrane manufacturers for their elements or, where
available, on the “stabilized salt rejection”.
As Tables 5.37, 5.38, and 5.39 in Annexes 5.A2, 5.A2.1, and 5.A3 show, there
are many membrane types and membrane elements available on the RO membrane

Table 5.5 Product and salt permeability coefficients of commercial seawater and brackish water
RO membranes at standard conditions
Product permeability (A value) Salt permeability (B value)
AE(St),V BNaCl, E(St)
m3/m2, h, bar
Type of membrane element [ 103] l/m2, h, bar m/h [ 105]
Seawater membranes
Spiral-wound composite
Normal 1.0–1.6 6.0–8.2
High rejection 0.9–1.1 3.5–5.60
High flux 1.2–1.9 6.2–11.5
Low fouling 1.1–1.8 8.3–10.2
Spiral-wound nanocomposite
High/highest rejection 1.0–1.4 4.1–5.7
High flux 2.2–2.3 11.6–11.7
Hollow-fibre CTA ~0.1 ~0.7
Brackish water membranes
Spiral-wound composite
Normal 3.3–3.9 13.2–14.8
High rejection 2.0–3.5 6.6–9.3
High flux/lower salt rejection 4.4–5.9 16.3–40.4
Spiral-wound nanocomposite
High rejection 3.3–3.7 10.5–12.5
High flux/lower salt rejection 5.5–6.5 17.5–45.0
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 407

market, which differ in the material and structure of their membrane separation layer,
the membrane area per element, and their performance characteristics. Calculation of
the AE(St),V and BNaCl,E(St) values or the permeability quotient Pi, which is equal to
AEðStÞ,V
BNaCl,EðStÞ , permits a comparison of the membrane-related performance characteristics
of different membrane elements. If the comparison reveals a higher Pi value, this is
an indication that the membrane has a low salt flux, i.e. a high salt rejection
capability.
As can be seen from Table 5.5, hollow-fibre CTA membranes have a much lower
permeability coefficient than polyamide TFC membranes. However, another crite-
rion for a membrane element’s production performance is its packing density, which
is the membrane area SE referred to the element’s volume VE in accordance with
Eq. (5.77).

SE
Pd ¼ : ð5:77Þ
VE
Pd ¼ Packing density [m2/m3]
SE ¼ membrane area of membrane element [m2]
VE ¼ geometric volume of element [m3]
A comparison of the packing density of polyamide spiral-wound elements with
hollow-fibre CTA elements is given in Table 5.6. This shows that a CTA hollow-
fibre element has a membrane area that is many times greater than that of a spiral-
wound module. As a result, despite its substantially lesser permeability, such a
membrane element has a production performance that is comparable with that of
an 8" polyamide spiral-wound element.
To compare various membrane elements regarding their element-specific perfor-
mance characteristics, the “relative salt transport value” (RSTV) finds application
(Eq. 5.78). The RSTV is a parameter used in practice and is calculated as the product
of an element’s product flux JPE(St) and its percentage salt passage SPNaCl(St),E, with
both of these based on standard conditions [52].

Table 5.6 Packing density of commercial RO membrane elements


Element size
Membrane area SE Diameter Length Packing density Pd
Membrane type m2 mm mm m2/m3
Spiral-wound composite
8" 34.4 201 1016 1067
35.3 201 1095
37.2 201 1154
40.9 201 1269
16" 148.6 402 1152
158.0 402 1225
163.5 402 1268
Hollow-fiber CTA 1200 280 2078 9400
408 5 Reverse Osmosis Membrane System: Core Process of SWRO

RSTV ¼ J PEðStÞ  SPNaClðStÞ,E : ð5:78Þ

RSTV ¼ relative salt transport value


JPE(St) ¼ product flux of element at standard conditions [l/m2, h]
SPNaCl(St),E ¼ salt passage rate of element at standard conditions [%]
An RSTV that shows up in the comparison as being lower indicates that the
membrane has favourable salt rejection.

Extrapolation from Standard Conditions to Design or Operation Conditions


If a membrane element’s performance data under standard test conditions are known,
by applying the relevant equations for water and salt transport, its performance data
can be determined under design or operating conditions that deviate from standard
conditions. This can be done either on the basis of the coefficients for water
permeability AE(St), V and salt permeability BiE(St) or with the help of correction
factors for the product flux fFC and the salt passage f SPCi . These correction factors
may be obtained by setting the mass transport equations at the required design
conditions in relation to those at standard conditions.
If the water permeability coefficient AE(St), V under standard conditions X-
(St) together with the correction factors for temperature fTCW(d ) and the reduction
of water permeability f FP,E,τMOpðdÞ with the membrane’s operating time τMOp are
known, the product flux J P,EðdÞ,V,τMOp and production performance of an element
F P,EðdÞ,τMOp under design conditions X(d ) are calculated with Eqs. (5.79), (5.80), and
(5.81).

J P,EðdÞ,V,τMOp ¼ AEðStÞ,V  f TCWðdÞ  f FP,E,τMOpðdÞ  NDPEðdÞ,τMOpðdÞ : ð5:79Þ

F P,EðdÞ,τMOp ¼ SE  J P,EðdÞ,V,τMOp : ð5:80Þ

F P,EðdÞ,τMOp
J P,EðdÞ,V,τMOp ¼ : ð5:81Þ
SE
fFC ¼ product flow correction factor []
X(d ) ¼ design conditions
X(d ), τMOp(d ) ¼ design conditions at operation time τMOp of membrane element
The correction factor fFC is derived from Eqs. (5.79) and (5.82) and the algorithm
for water transport under standard conditions. In accordance with Eq. (5.82a), it is
calculated from:

• the correction factor for the influence of temperature on water permeability, fTCW
(d )
• the correction factor for the influence of the membrane’s operating time on water
permeability, f FW,E,τMOpðdÞ
• the ratio of the net driving pressure under design conditions (Eq. 5.82b) to its
value under standard conditions (Eq. 5.82c), NDPEðdÞ,τMOpðdÞ .
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 409

F P,E ðdÞ,τMOp J P,EðdÞ,V,τMOp


f FC ¼ ¼ : ð5:82Þ
F P,EðStÞ J P,EðStÞ,V

NDPEðdÞ,τMOpðdÞ
f FC ¼ f TCWðdÞ  f FP,E,τMOpðdÞ  : ð5:82aÞ
NDPEðStÞ

However, for both these calculation approaches, the required feed pressure to the
membrane element under design conditions pF,EðdÞ,τMOpðdÞ has to be known so as to be
able to determine the NDPEðdÞ,τMOpðdÞ at design conditions in accordance with
Eq. (5.82b). The NPDE(St) is calculated from the standard test conditions using
Eq. (5.82c).

ΔpFC,EðdÞ
NDPEðdÞ,τMOpðdÞ ¼ pF,EðdÞ,τMOpðdÞ  pP,EðdÞ 
2
 
 π F ðdÞ,NaCl  CFEðdÞ  βEðdÞ  π P,EðdÞ : ð5:82bÞ

ΔpFC,EðStÞ  
NDPEðStÞ ¼ pF,EðStÞ   π FðStÞ  CFEðStÞ  βEðStÞ  π P,EðStÞ : ð5:82cÞ
2
Differing values for product recovery YE(d ),(St) under standard and under design
conditions impact the membrane’s water permeability by way of the value of the
respective average concentration factor CFEðdÞ,ðStÞ (Eqs. 5.49, 5.50, and 5.51) or that
of the average concentration polarization factor βEðdÞ,ðStÞ (Eq. 5.55) when calculating
the respective net driving pressure NDP(d ),(St) (Eqs. 5.82b and 5.82c).
When calculating the net driving pressure NDPEðdÞ,τMOpðdÞ under design conditions,
the osmotic pressure of the feed to the element π F(d ) is not referred to the feed salt
concentration under design conditions cF(d ), but to its sodium chloride equivalent
cNaCl,equiv. This value is calculated from the actual osmotic pressure of the feed
π F(d ) in accordance with Eq. (5.82d). By applying Eq. (5.82e), the osmotic pressure
associated with this π F(d ), NaCl may be determined from the sodium chloride equiva-
lent of the feed salt content cNaCl, equi and this then provides the basis for calculating
the net driving pressure under design conditions NDPE(d ).

π FðdÞ  ρF ðdÞ 1000  SF ðdÞ


cNaCl,equi ¼ 3:51195  102    : ð5:82dÞ
∅F ðdÞ  273:15 þ t ðdÞ 1000

1 1000
π FðdÞ,NaCl ¼ 2:8474  103  ð273:15 þ t Þ  
ρFðdÞ 1000  SF ðdÞ
 ∅F ðdÞ  cNaCl,equi : ð5:82eÞ

cNaCl, equi ¼ sodium chloride equivalent of salt content [g/l]


ρF(d ) ¼ density of feed at design conditions [kg/l]
410 5 Reverse Osmosis Membrane System: Core Process of SWRO

SF(d ) ¼ salinity of feed [g/kg]


π F(d ), NaCl ¼ osmotic pressure at design conditions based on the NaCl-equivalent
of feed concentration ciFEðdÞ [bar]
∅F(d ) ¼ osmotic coefficient of feed at design conditions []
The two equations for calculating the sodium chloride equivalent and the osmotic
pressure associated with this (Eqs. 5.82d and 5.82e) are obtained by appropriate
transformation of Eq. (3.115a) in Sect. 3.2.3.3.
The feed pressure needed for the membrane element under design conditions
pF,EðdÞ,τMOpðdÞ is then determined from the specified production performance of the
element F P,EðdÞ,τMOp in accordance with Eq. (5.83) and, with this, then the
corresponding NDPE(d ) under design conditions is calculated with Eq. (5.82b).

F P,EðdÞ,τMOp ΔpFC,EðdÞ
pF,EðdÞ,τMOpðdÞ ¼ þ pP,EðdÞ þ
SE  AEðStÞ,V  f TCWðdÞ  f FP,E,τMOpðdÞ 2
 
þ π F ðdÞ,NaCl  CFEðdÞ  βEðdÞ  π P,EðdÞ : ð5:83Þ

pF,EðdÞ,τMOpðdÞ = feed pressure of membrane element at design conditions and for an


operation time τMOp(d ) [bar]
The feed pressure of the membrane element at design conditions pF,EðdÞ,τMOpðdÞ is
changed in line with the ratio of the specific product flux at design conditions
J PS,EðdÞ,V,τMOp to that under standard conditions JP, S, E(St), V in accordance with
Eq. (5.83a).

J PS,EðdÞ,V,τMOp NDPEðStÞ
pF,EðdÞ,τMOpðdÞ ¼  þ pP,EðdÞ
J PS,EðStÞ,V f TCWðdÞ  f FP,E,τMOpðdÞ
ΔpFC,EðdÞ  
þ þ π FðdÞ,NaCl  CFEðdÞ  βEðdÞ  π P,EðdÞ : ð5:83aÞ
2
For seawater desalination, it may be assumed that pP,EðdÞ pF,EðdÞ,τMOpðdÞ ,
π P,EðdÞ pF,EðdÞ,τMOpðdÞ and ΔpFC,EðdÞ pF,EðdÞ,τMOpðdÞ . Then Eq. (5.83a) is accord-
ingly simplified to Eq. (5.83b).

J PS,EðdÞ,V,τMOp NDPEðStÞ
pF,EðdÞ,τMOpðdÞ ¼  þ π F ðdÞ,NaCl
J PS,EðStÞ,V f TCWðdÞ  f FP,E,τMOpðdÞ
 CFEðdÞ  βEðdÞ : ð5:83bÞ

An important parameter for dimensioning membrane elements is the element’s


specific product flux J PS,EðdÞ,V,τMOp . This corresponds to the product flux J P,EðdÞ,V,τMOp ,
but is measured in the units [l/m2, h] or also [l mh]. This is calculated from the
product flux or the production performance with Eq. (5.83c).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 411

F P,EðdÞ,τMOp
J PS,EðdÞðStÞ,V,τMOp ¼ J P,EðdÞðStÞ,V,τMOp  1000 ¼  1000: ð5:83cÞ
SE
J PS,EðdÞðStÞ,V,τMOp ¼ specific product flux of membrane element at design/standard
conditions [l/m2, h]
After calculating pF,EðdÞ,τMOpðdÞ , then with Eq. (5.82b) the value of the
NDPEðdÞ,τMOpðdÞ can be calculated and from this, with Eq. (5.82a), the correction factor
fFC for the membrane element’s product flux under design conditions can be
determined.
Regarding the element’s production performance F P,EðdÞ,τMOp and its product flux
J P,EðdÞ,V,τMOp , these two parameters are calculated with the correction factor fFC, as
shown in Eqs. (5.84) and (5.84a).

F P,EðdÞ,τMOp ¼ F P,EðStÞ  f FC
NDPEðdÞ,τMOpðdÞ
¼ F P,EðStÞ  f TCWðdÞ  f FP,E,τMOpðdÞ  : ð5:84Þ
NDPEðStÞ

J P,EðdÞ,V,τMOp ¼ J P,EðStÞ,V  f FC
NDPEðdÞ,τMOpðdÞ
¼ J P,EðStÞ  f TCWðdÞ  f FP,E,τMOpðdÞ  : ð5:84aÞ
NDPEðStÞ

Like for the water transport in a membrane element, it is also possible to


determine its parameters for salt transport and salt rejection either with the aid of
the membrane-specific water and salt permeability coefficients AE(St), V and BiE(St), or
by using the correction factor for the salt passage of the respective dissolved
components f SPCi . Equation (5.85) shows the calculation for the product concentra-
tion ciP,Eðd,τMOpðdÞ Þ of a salt component i under design conditions from its feed
concentration ciFEðdÞ to the membrane element with the help of the salt permeability
coefficient BiE(St) and the water permeability coefficient AE(St), V under standard
conditions.

BiEðStÞ  f TCSðdÞ  f FSi,E,τMOp  ciFEðdÞ  CFEðdÞ  βEðdÞ


ciP,Eðd,τMOpðdÞ Þ ¼ : ð5:85Þ
AEðStÞ,V  f TCWðdÞ  f FP,E , τMOpðdÞ  NDPEðdÞ,τMOpðdÞ

The corresponding salt passage SPiE ðdÞ,τMOpðdÞ of the component i results as shown
by Eq. (5.85a). To determine the product concentration of component i on the basis
of the salt passage, see Eq. (5.85b).
412 5 Reverse Osmosis Membrane System: Core Process of SWRO

 1
AEðStÞ,V  f TCWðdÞ  f FP,E,τMOpðdÞ  NDPEðdÞ,τMOpðdÞ
SPiE ðdÞ,τMOpðdÞ ¼ 1þ :
BiEðStÞ  f TCSðdÞ  f FSi,E,τMOp
ð5:85aÞ

SPiEðdÞ,τMOpðdÞ ¼ salt passage factor of component i at design conditions for the


membrane element []

ciP,EðdÞ,τMOpðdÞ ¼ ciFEðdÞ  CFiEðdÞ  βEðdÞ  SPiEðdÞ , τMOpðdÞ : ð5:85bÞ

Regarding the derivation of the above equations for salt passage and product
concentration of the component i from the membrane mass transport equations,
reference is made to Sect. 5.1.5.1.2, Eqs. (5.16), (5.20e), 5.20f, (5.21), (5.22), and
(5.23).
Like the correction factor fFC for the product flux of a membrane element, the salt
passage correction factor f SPCi for the component i is derived from the ratio of the
mass transport equations at the element’s design conditions to those under standard
conditions (Eq. 5.86). As shown by Eq. (5.86a) for the factor f SPCi , its value
depends on the ratio of the design to the standard conditions, i.e. of the:

• net driving pressures NDPEðdÞ,τMOpðdÞ and NDPE(St)


• concentration factors CFiEðdÞ CFiEðStÞ
• concentration polarization factors βEðdÞ and βEðStÞ
• as well as the quotients of
– the temperature correction factor for water permeability fTCW(d ) and the salt
permeability of the membrane fTCS(d ) under design conditions
– the correction factors that are dependent on the membrane operating time for
how the water and salt permeability f FW,E,τMOpðdÞ and f FSi,E,τMOp change.

SPiE ðdÞ,τMOpðdÞ
f SPCi ¼ : ð5:86Þ
SPiEðStÞ

f TCSðdÞ  f FSi,E,τMOp CFiEðdÞ  βEðdÞ NDPEðStÞ


f SPCi ¼   : ð5:86aÞ
f TCWðdÞ  f FP,E,τMOpðdÞ CFiEðStÞ  βEðStÞ NDPEðdÞ,τMOpðdÞ

f SPCi ¼ salt passage correction factor of component i []


The salt passage of a component i under design conditions SPiMEðdÞ,τMOpðdÞ is
calculated from the salt passage under standard conditions SPiME(St) and then by
multiplying the standard salt passage with the salt passage correction factor f SPCi
(Eq. 5.86b).
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 413

SPiEðdÞ,τMOpðdÞ ¼ SPiEðStÞ  f SPCi : ð5:86bÞ

SPiE(St) ¼ membrane salt passage factor of component i at standard test conditions


[]

5.1.5.2.4 Salt Passage and Product Water Composition


To determine the concentrations of the individual ion types of the water to be treated
in the permeate of a membrane element and thus also to ascertain a corresponding
product water analysis, either the specific permeability coefficients for each of the
ions BiE(St) or their salt passages SPiME(St) have to be known, with both under
standard conditions.
The respective concentration of a single component ciP,EðdÞ,τMOpðdÞ in the permeate
is calculated by using the water and salt permeability coefficients AE(St), V and BiE(St)
under standard conditions as shown above with Eqs. (5.85a) and (5.85b) or by
calculating the product composition with the help of the salt passage correction
factor f SPCi corresponding to Eqs. (5.86a) and (5.86b) as well as (5.85b).
The total salt content of the element’s product water under design conditions
cP,EðdÞ,τMOpðdÞ or the TDSP,EðdÞ,τMOpðdÞ then results from the sum of the concentrations of
the individual components ciP,EðdÞ,τMOpðdÞ (Eqs. 5.87 and 5.87a).
X
cP,EðdÞ,τMOpðdÞ ¼ ciP,EðdÞ,τMOpðdÞ : ð5:87Þ
i

TDSP,EðdÞ,τMOpðdÞ ¼ 1000  cP,EðdÞ,τMOpðdÞ : ð5:87aÞ

cP,EðdÞ,τMOpðdÞ ¼ total salt content in the membrane element product at design


conditions [kg/m3; g/l]
TDSP,EðdÞ,τMOpðdÞ ¼ total dissolved solids content in the membrane element product
at design conditions [mg/l]
Following the addition of all of the calculated ion concentrations in the product
water, to obtain a consistent overall analysis, using the ion equivalents of each of the
component concentrations as described in Sects. 3.1 and 3.1.1, and with the aid of
Eqs. (3.17), (3.18), (3.19), and (3.20), a test of electrical neutrality has to be
conducted with the results of this then used to reconcile the ion concentrations by
raising the cation or anion equivalents to attain electrical neutrality in accordance
with Eqs. (3.21) and (3.22).
Indicative values of salt passage and salt rejection for cations and anions of
seawater at standard test conditions are compiled in Table 5.7 for an RO membrane
for seawater desalination with an NaCl rejection rate of 99.8% under standard
conditions. Such data serve as a basis for extrapolating the salt passages of individual
ions at standard conditions to the corresponding values for the elements under the
design conditions of a membrane system.
414 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.7 Indicative Salt passage Salt rejection


values of ionic salt passage SPi(St) Ri(St)
and rejection of a PA com-
Component % %
posite high rejection sea-
water membrane Cations
(RNaCl ¼ 99.8%) at standard Calcium Ca2+ 0.05 99.95
conditions Magnesium Mg2+ 0.05 99.95
Sodium Na+ 0.23 99.77
Potassium K+ 0.30 99.70
Barium Ba2+ 0.05 99.95
Strontium Sr2+ 0.05 99.95
Anions
Chloride Cl 0.23 99.77
Sulphate SO42 0.06 99.94
Nitratea NO3 0.35 99.65
Bicarbonatea HCO3 0.35 99.65
Carbonatea CO32 0.05 99.95
Fluoridea F 0.35 99.65
Bromide Br 0.27 99.74
Borona,b 5.00 95.00
Carbon dioxide CO2 100.00 0
TDS 0.20 99.80
a
Rejection is pH dependent
b
At pH ¼ 8

Bromide Salt Passage


During investigations undertaken by a membrane manufacturer at plants equipped
with seawater desalination membranes with a high salt rejection rate, it became
apparent that there was a reproducible relationship between the salt passage of
bromide and that of chloride. On this basis, it is possible to determine the bromide
concentration in the product of a plant cBr ,P,EðdÞ,τMOpðdÞ from its feed concentration
cBr ,F,EðdÞ and the existing chloride salt passage (Eqs. 5.88, 5.88a, 5.88b, and
5.88c) [53].

cCl P,EðdÞ,τMOpðdÞ
cBr ,P,EðdÞ,τMOpðdÞ ¼ 1:15  cBr ,F,EðdÞ  : ð5:88Þ
cCl ,F,EðdÞ

cCl P,EðdÞ,τMOpðdÞ ¼ cCl ,F,EðdÞ  CFEðdÞ  βEðdÞ  SPCl ,MEðdÞ,τMOpðdÞ : ð5:88aÞ

cBr ,P,EðdÞ,τMOpðdÞ ¼ 1:15  cBr ,F,EðdÞ


cCl ,F,EðdÞ  CFEðdÞ  βEðdÞ  SPCl ,MEðdÞ,τMOpðdÞ
 : ð5:88bÞ
cCl ,F,EðdÞ
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 415

cBr ,P,EðdÞ,τMOpðdÞ ¼ 1:15  cBr ,F,EðdÞ  CFEðdÞ  βEðdÞ  SPCl ,MEðdÞ,τMOpðdÞ : ð5:88cÞ

cCl ,F,EðdÞ = chloride feed concentration at design conditions [mg/l]


cBr ,F,EðdÞ = bromide feed concentration at design conditions [mg/l]
cCl ,P,EðdÞ,τMOpðdÞ ¼ chloride concentration in product at design conditions and for a
membrane operation time τMOp [mg/l]
cCl ,P,EðdÞ,τMOpðdÞ ¼ chloride concentration in product at design conditions and for a
membrane operation time τMOp [mg/l]
SPCl ,MEðdÞ,τMOpðdÞ ¼ intrinsic salt passage factor of chloride at design conditions
and for a membrane operation time τMOp []
The above relationships apply for a bromide concentration in seawater of up to
100 mg/l. If this value is exceeded, the bromide feed concentration cBr ,F,EðdÞ should
be added to the chloride feed concentration cCl ,F,EðdÞ as chloride equivalent.

Dependence of Salt Passage of the CO2/Bicarbonate/Carbonate Equilibrium


Components on pH
The salt rejection of various anions or anion groups is strongly dependent on the pH
of the feed to the membrane element.
In particular, the degree of salt passage of the anions bicarbonate HCO3 and
carbonate CO32 that are responsible for the CO2/bicarbonate/carbonate equilibrium
in seawater as well as the free carbonic acid H2CO3 and carbon dioxide CO2 is
greatly dependent on pH (Eqs. 5.89 and 5.89a).
   
½Hþ   HCO ½10pH  HCO
  3 ¼   3
¼ K1 ð5:89Þ
CO2 CO2
   
½Hþ   CO2 ½10pH  CO2
  3
¼   3
¼ K2 ð5:89aÞ
HCO 3 HCO 3

K*1 ¼ 1st stoichiometric dissociation constant of carbonic acid [mol/kg]


K*2 ¼ 2nd stoichiometric dissociation constant of carbonic acid [mol/kg]
If prior to its desalination acid is added to the seawater to suppress scaling, these
equilibrium proportions are shifted. The pH of the pretreated seawater which is fed to
the RO membrane is decisive for the new ratio of the three dissolved components to
each other. According to the pH that is set, the concentrations of HCO3, CO32, and
CO2 in the feed to the membrane element must therefore be recalculated based on
their original concentrations in seawater.
The new proportions and the resulting concentration of each component may be
recalculated either using carbonate alkalinity ACF (Eqs. T5.8.1a–T5.8.1g in
Table 5.8) or dissolved inorganic carbon CTF (Eqs. T5.8.2a–T5.8.2g in Table 5.8).
But strictly speaking, the equations set out in Table 5.8 apply only for a pH range of
around 6 to 8.5. For pH values that are outside this range, more complex algorithms
must be used that also take into account additional influences on the CO2/
416 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.8 Equations for bicarbonate/carbonate/CO2 equilibrium distribution factors and equilib-
rium components concentration calculation
Parameters
Result to be Equation
Settings available calculated Algorithm no.
   2     2 
HCO3 F , CO3 F
AC F HCO3 F þ2 CO3 F T5.8.1a
AC F , [H+]F ½CO2 F ½H þ F
2
T5.8.1b
AC F ¼ F CO2 K1 ð½H þ F þ2 K2 Þ

½HCO3 F ½H þ F T5.8.1c
AC F ¼ F HCO3 ½H þ F þ2 K 2

½CO2
3 F
K2 T5.8.1d
AC F ¼ F CO3 ½H þ F þ2 K 2

AC F , F CO2 ,ρF cCO2 ,F,EðdÞ AC F 44, 0 ρF F CO2 T5.8.1e


AC F , F HCO3 ,ρF cHCO3 ,F,EðdÞ AC F 61, 0 ρF F HCO3 T5.8.1f
AC F , F CO3 ,ρF cCO3 ,F,EðdÞ AC F 60, 0 ρF F CO3 T5.8.1g
       2 
HCO , CT F CO2 þ HCO 3 þ CO3
T5.8.2a
 23 F  
CO3 F , CO2 F
C T F , [H+]F ½CO2 F ½H þ F
2
T5.8.2b
CT F ¼ F CO2 ½H þ F þK 1 ½H þ F þK 1 K 2
2

½HCO3 F K 1 ½H þ F T5.8.2c
CT F ¼ F HCO3 ½H þ F þK 1 ½H þ F þK 1 K 2
2

½CO2
3 F
K1 K2 T5.8.2d
CT F ¼ F CO3 ½H þ F þK 1 ½H þ F þK 1 K 2
2

AC F , F CO2 ,ρF cCO2 ,F,EðdÞ C T F 44, 0 ρF F CO2 T5.8.2e


AC F , F HCO3 ,ρF cHCO3 ,F,EðdÞ C T F 61, 0 ρF F HCO3 T5.8.2f
AC F , F CO3 ,ρF cCO3 ,F,EðdÞ C T F 60, 0 ρF F CO3 T5.8.2g
 

ACF = carbonate alkalinity of feed [mol/kg] HCO3 F = bicarbonate feed concentration
CTF = DIC = total dissolved inorganic carbon
content of feed [mol/kg] [mol/kg]
CO2
3 F = carbonate feed concentration
ρF = density of feed [kg/m3] [mol/kg]
F CO2 , FHCO3, , F CO3 = distribution rate factors  
CO2 F = carbon dioxide feed concentration
[mol/kg]
[H+]F = hydrogen ion feed concentration
[mol/kg]
cCO2 ,F,EðdÞ = carbon dioxide concentration in feed at ½H þ F =pHF [mg/l]
cHCO3 ,F,EðdÞ = bicarbonate concentration in feed at ½H þ F =pHF [mg/l]
cCO3 ,F,EðdÞ = carbonate concentration in feed at ½Hþ F =pHF [mg/l]
K*1 = 1st stoichiometric dissociation constant of carbonic acid [mol/kg]
K*2 = 2nd stoichiometric dissociation constant of carbonic acid [mol/kg]

bicarbonate/carbonate equilibrium in seawater. More detailed information on the


CO2/bicarbonate/carbonate equilibrium in seawater, the influence of other seawater
components such as boric acid compounds on the equilibrium, and algorithms to
calculate the equilibrium concentration proportions with the resulting pH can be
found under Sects. 3.2.4 and 3.2.4.1, Table 3.14. This section also contains an
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 417

equation for calculating the first dissociation constant K*1 and the second dissocia-
tion constant for carbonic acid K*2, both of which are needed for determining the
distribution factors FCO2 , FHCO3 , and FCO3 (Eqs. 3.123 and 3.123a). The pK*1 and
pK*2 values with their dependency on salinity and temperature can also be taken
from the graphs of Figs. 3.36 and 3.37 of this section. The range in which this
algorithm is valid, though, as stated there extends only up to a salinity of 50 g/kg. For
higher salinities of up to 70–80 g/kg, like the concentrations encountered in mem-
brane desalination of seawater, this equation system may still be used if certain
inaccuracies in the salt concentration range can be tolerated, in particular for rule-of-
thumb calculations for practical and not for scientific purposes. If this is not the case,
the distribution factors of the CO2/bicarbonate/carbonate equilibrium at higher
salinities can be calculated with the Pitzer equations (see Sect. 3.2.3.2.1,
Eqs. 3.102 and 3.103–3.103j) with the help of the PHREEQC software and its Pitzer
database pitzer.dat (see Sect. 3.2.3.2.1).
As Table 5.7 shows, there are marked differences between the salt passages of
bicarbonate and carbonate. Virtually, no CO2 is intercepted by an RO membrane.
But this also means that, with subsequent desalination, for example at each mem-
brane element linked in series of a membrane module, the proportions of the three
equilibrium components with respect to each other continue to change. Due to the
transfer of most of the CO2 into the permeate, the CO2 in the concentrate is depleted,
so there the pH rises while it decreases in the permeate. At the same time, in the
concentrate the ratio HCO3/CO32 shifts to an increased carbonate content.

pH Dependency of Salt Passage Rate of Boron Compounds and Boron Salt


Passage
Boron is present in seawater as a mixture of boric acid and borate. The shares of
these two components in the boric acid/borate equilibrium are fixed by the dissocia-
tion constant of the boric acid K B and the concentration of H+ ions, i.e. the seawater’s
pH (Eq. 5.90).
 
BðOHÞ ½Hþ 
 4  ¼ KB ð5:90Þ
BðOHÞ3

K*B ¼ stoichiometric dissociation constant of boric acid [mol/kg]


The corresponding ratios of the respective boron components to the total boron
content, i.e. the factors F BðOHÞ3 and F BðOHÞ4 of the distribution relationship, are
calculated with Eqs. (5.91) and (5.91a). The total boron content is the sum of the
respective concentrations of boric acid and borate (Eq. 5.92).
 
BðOHÞ4 KB
F
¼ F BðOHÞ4  ¼ ð5:91Þ
BTF ½Hþ F þ K B
418 5 Reverse Osmosis Membrane System: Core Process of SWRO

 
BðOHÞ3 F K ½Hþ F
¼ FBðOHÞ3 ¼ 1  þ B ¼ þ ð5:91aÞ
BTF ½H F þ K B ½H F þ K B
X    
BF ¼ BðOHÞ 4 F þ BðOHÞ3 F ð5:92Þ

∑BF ¼ total boron content [mol/kg]


[B(OH)
 3]F ¼ boric acid feed concentration [mol/kg]
BðOHÞ 4 F ¼ borate feed concentration [mol/kg]
FBðOHÞ3 , FBðOHÞ4 , ¼ distribution rate factors of boric acid/borate equilibrium
components [–]
The value of the dissociation constant K*B of boric acid depends on the seawater’s
salt content and its temperature. An equation for calculating ln K*B is to be found in
Sect. 3.2.4.2, Eq. (3.139) and the associated pK*B values there in Fig. 3.41 as well as
in Annex 3.A2—Table 3.33 of Chap. 3. Also, the factors FBðOHÞ3 and FBðOHÞ4 are
presented as graphs in Sect. 3.2.4.2 in Figs. 3.42 and 3.43, showing their dependency
on pH and salinity as well as on pH and temperature.
As noted above for the calcium/carbonic acid equilibrium in seawater, the limited
validity of Eq. (3.139) for salinity in the range of 5–45 g/kg has to be taken into
consideration.
Besides the Dickinson Equation (Eq. 3.139), publications on the boron perme-
ability of RO membranes include other algorithms for calculating the dissociation
constant K*B as functions of temperature and salt content. One of these is the
equation put forward by Messmer [54], (Eq. 5.93), whose validity with regard to
salt content lies within a range for ionic strength of 0 to 1. This may be applied to
both brackish water and seawater for calculating the dissociation constant of boric
acid.

1573:21
log K B,OH ¼ þ 28:6059 þ 0:012078  T  13:2258  log T
T
þ ð0:3250  0:00033  T Þ  I  0:0912  I 3=2 : ð5:93Þ

Validity I ¼ 0  1 m, T ¼ 273–573 K
 
BðOHÞ
 4
 ¼ K B,OH ð5:93aÞ
½OH  BðOHÞ3

pK B ¼ pK W  pK B,OH ð5:93bÞ

K*W ¼ stoichiometric ion product of water [mol/kg]


T ¼ temperature [K]
The equation is based on the dissociation equilibrium of boric acid with the
hydroxyl concentration, i.e. the pOH value, according to Eq. (5.93a). From the so
calculated dissociation coefficient K B,OH or pK B,OH , the coefficient referred to the
[H+] concentration K B is calculated with Eq. (5.93b). The value of the
Table 5.9 Seawater RO polyamide membranes: indicative values of boric acid and borate permeability, membrane temperature coefficients, and reflection
coefficients [54]
BBðOHÞ3 ,Eð25 CÞ σ BðOHÞ3 A BðOHÞ4  ,Eð25 CÞ σ BðOHÞ4  b
 1  1
Type of membrane m/h – C m/h – C
SW high boron rejection 9.8  104–13.3  104 0.983–0.996 0.051–0.068 1.8  104–2.6  104 0.997–0.999 0.033–0.069
SW high flux 1.3  103–1.7  103 0.995–0.997 2.3  104–2.7  104 0.997–0.999
SW lower rejection 2.0  103 0.975 3.2  104 0.998
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application
419
420 5 Reverse Osmosis Membrane System: Core Process of SWRO

stoichiometric ion product K*W or the associated pKW value for water as functions of
salinity and temperature is calculated using Eqs. (3.124) and (3.125) (Sect. 3.2.4.1).
With the various algorithms for calculating the dissociation coefficient of boric
acid found in the literature, marked differences are obtained both in the calculated
values and how the coefficients so obtained vary with salt content and temperature.
As a consequence, depending on the chosen manner of calculation of the boric acid
dissociation coefficient, also the results for the pH dependency of the membranes’
boron permeability differ, as shown in the following.
The distribution factors of the boric acid equilibrium, in particular for concentrate
salt contents of a seawater RO plant in the range of 70 g/l and more but naturally also
across the complete bandwidth of brackish water up to seawater concentrate, can
also be calculated by thermodynamic modelling on the basis of the Pitzer equations
and with the help of PHREEQC software.
RO desalination membranes exhibit differing permeabilities for boric acid and
borate. In this respect, the rejection of the undissociated boric acid molecule is
substantially worse than it is for the negatively charged borate ion. The ratio of the
permeabilities is specific for the membranes of each manufacturer. Table 5.9 shows a
comparison of indicative values of the permeabilities of boric acid and borate at
25  C for various types of TFC polyamide seawater RO membranes.

The total boron permeability of a desalination membrane BB, E(St) under standard
conditions is made up of:

• the membrane-specific standard permeabilities for boric acid and borate


• the concentration ratio of these two components which in turn depends on the
seawater’s pH, salt content, and temperature (Eq. 5.94).

The temperature dependency of the permeability results from a membrane-


specific temperature correction factor for the respective boron component
(Eqs. 5.95, 5.95a, 5.95b, and 5.95c) [55, 56].

BB,EðStÞ ¼ F BðOHÞ3  BBðOHÞ3 ,EðStÞ þ F BðOHÞ4   BBðOHÞ4  ,EðStÞ : ð5:94Þ

BBðOHÞ3 ,EðdÞ ¼ BBðOHÞ3 ,EðStÞ  f TC,BðOHÞ3ðdÞ ð5:95Þ

f TC,BðOHÞ3ðdÞ ¼ eaðtðdÞ tðStÞ Þ ð5:95aÞ

BBðOHÞ4  ,EðdÞ ¼ BBðOHÞ4  ,EðStÞ  f TC,BðOHÞ4  ðdÞ ð5:95bÞ

f TC,BðOHÞ4  ðdÞ ¼ ebðtðdÞ tðStÞ Þ : ð5:95cÞ

BB, E(St) ¼ boron permeability at standard conditions [m/h]


BðOHÞ3 ,EðStÞ ¼ boric acid permeability at standard conditions [m/h]
BðOHÞ4  ,EðStÞ ¼ borate permeability at standard conditions [m/h]
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 421

f TC,BðOHÞ3ðdÞ ¼ temperature correction factor for boric acid permeability []


f TC,BðOHÞ4  ðdÞ ¼ temperature correction factor for borate permeability [])
a ¼ membrane specific coefficient []
b ¼ membrane specific coefficient []
BB, E(d ) ¼ boron permeability at design conditions [m/h]
BðOHÞ3 ,EðdÞ ¼ boric acid permeability at design conditions [m/h]
BðOHÞ4  ,EðdÞ ¼ borate permeability at design conditions [m/h]
The total boron permeability of a membrane element under design conditions is
calculated from the boric acid/borate distribution corresponding to the salt content
and the temperature under these conditions and the respective temperature-corrected
permeability of the two components according to Eq. (5.95d) as well as Eqs. (5.95),
(5.95a), (5.95b), and (5.95c).

BB,EðdÞ ¼ F BðOHÞ3 ðdÞ  BBðOHÞ3 ,EðdÞ þ F BðOHÞ4  ,EðdÞ  BBðOHÞ4  ,EðdÞ ð5:95dÞ

From the total boron permeability of a membrane element under standard


conditions calculated in accordance with Eqs. (5.94) or (5.95d) and knowing the
element’s corresponding product flux, the boron rejection rate RB, E(d ) can be
calculated with Eqs. (5.96), (5.96a), and (5.96b). Additionally, the reflection coeffi-
cient σB has to be known. Because the boron rejection rate of an RO membrane is
appreciably lower that its rejection rate for other salt components, it is no longer
possible to calculate this using the solution diffusion model. Instead, coupling of
volume and salt transport has to be incorporated in line with the Spiegler-Kedem
model. This is done with the help of the reflection coefficient σ B, which takes into
account the degree of this coupling [44]. Table 5.9 also contains values of this
coefficient for the two components of the boric acid equilibrium. The reflection
coefficient σ B(d ) resulting for the particular pH conditions is calculated with
Eq. (5.96b) from the respective coefficients σ BðOHÞ3 and σ BðOHÞ4  of the components
as well as the distribution coefficients F BðOHÞ3 and F BðOHÞ4 of boric acid and borate.
 
σ BðdÞ  1  f BðdÞ
RB,EðdÞ ¼ : ð5:96Þ
1  σ BðdÞ  f BðdÞ


ð
J PE ðd Þ  1σ BðdÞ Þ
f BðdÞ = e BB,EðdÞ
ð5:96aÞ

σ BðdÞ ¼ F BðOHÞ3ðdÞ  σ BðOHÞ3 þ F BðOHÞ4  ðdÞ  σ BðOHÞ4  ð5:96bÞ

RB, E(d ) = boron rejection rate []


σ B(d ) ¼ reflection coefficient at design conditions []
In addition to the parameters of the standard test conditions listed in Table 5.4,
boron rejection at standard conditions is referred by the membrane manufacturers to
a boron content of 5 mg B/l in the test solution and a specific pH. For seawater
polyamide membranes, this would normally be a pH of 8 and, depending on
422 5 Reverse Osmosis Membrane System: Core Process of SWRO

membrane type and manufacturer, the standard boron rejection rate would lie
between 90% and 96% (Annexes 5.A2, 5.A2.1—Tables 5.37 and 5.38). For brackish
water polyamide membranes, the standard rejection rate is specified for a pH of
10 and it is then between 93% and 96% (Annexe 5.A3—Table 5.39).
As a membrane element ages with time τMOp(d ), its boron permeability increases
which means that its boron rejection capacity drops. Like for the other salt content
components of the seawater, from the value of the coefficient at design conditions
BB, E(d ) multiplied by a membrane correction factor f FB,E,τMOp , there results for the
permeability coefficient of boron a specific membrane lifetime BB, E(d ),τMOp(d ) due to
the increase of its boron permeability with time (Eq. 5.97). The boron rejection of a
membrane element after a specified design operating time is given by Eqs. (5.97a)
and (5.97b).

BB,EðdÞ, τMOpðdÞ ¼ BB,EðdÞ  f FB,E,τMOp ð5:97Þ


 
σ BðdÞ  1  f BðdÞ,τMOpðdÞ
RB,EðdÞ, , τMOpðdÞ ¼ ð5:97aÞ
1  σ BðdÞ  f BðdÞ, τMOpðdÞ

J PE ðd Þ,τ
MOpðd Þ Þ
ð
 1σ Bðd Þ Þ
 BB,E ðd Þ,τ
f BðdÞ,τMOpðdÞ ¼ e MOpðd Þ Þ ð5:97bÞ

f FB,E,τMOp ¼ boron permeability correction factor of element at the element’s


operation time τMOp []
BB,EðdÞ,τMOpðdÞ ¼ boron permeability at design conditions and a membrane opera-
tion time τMOp [m/h]
BðOHÞ3 ,EðdÞ,τMOpðdÞ Þ ¼ boric acid permeability at design conditions and a membrane
operation time τMOp [m/h]
BðOHÞ4  ,EðdÞ,τMOpðdÞ Þ ¼ borate permeability at design conditions and a membrane
operation time τMOp [m/h]
As described above, by elevating the pH the share of borate in the total boron
content increases and, due to the lower permeability of B(OH)4, the total boron
permeability of a membrane increases so its boron rejection rate can be improved.
Figure 5.25 shows how pH influences the total boron permeability of a polyamide
seawater membrane with a higher boron rejection for a salt concentration as NaCl of
32,000 mg/l under standard test conditions.
This graph also shows the substantial influence of temperature on a membrane’s
boron rejection, particularly in the low or neutral pH range. If at a pH of 7.5, the
seawater’s temperature increases from 20  C to 30  C, the membrane’s boron
permeability approximately doubles while for a rise to 40  C it is around 3.5 times
as high. For a pH of 9 and under the same conditions, however, the permeability
increases by 2.7 times at the most and is at a substantially lower permeability level.
The boron permeability of a brackish water membrane that finds application in the
second pass of an SWRO plant for further post-desalination and boron reduction is
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 423

Boron permeability BB
[m/h]
3.5E-03

3.0E-03 RO seawater membrane - High boron rejection

2.5E-03

2.0E-03

1.5E-03 Temperature [°C)

1.0E-03
40

5.0E-04 30
25
20

0.0E+00
6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5
PH

Fig. 5.25 Seawater RO polyamide membrane: dependence of boron permeability on pH

Boron permeability BB
[m/h]

RO brackish water membrane - High boron rejection


Temperature [°C)
5.E-02
40

4.E-02

3.E-02

30
2.E-02

25

1.E-02 20

1.E-03
5.0 6.0 7.0 8.0 9.0 10.0 11.0
pH

Fig. 5.26 Brackish water RO polyamide membrane: dependence of boron permeability on pH

similarly dependent on pH and temperature (Fig. 5.26). The salt content corresponds
to the standard test conditions for brackish water membranes of 2000 mg/l NaCl. The
calculation of the distribution coefficients of boric acid and bromate for the diagram
424 5 Reverse Osmosis Membrane System: Core Process of SWRO

in Fig. 5.26 is based on boric acid dissociation coefficients K*B calculated with the
algorithm according to Eqs. (5.93), (5.93a), and (5.93b).
As the pH/temperature curves for the two membrane types show, by selecting an
appropriate pH range for both the seawater desalination unit and for post-desalination
of an RO system, the membrane elements’ boron rejection can be optimized
accordingly. For a seawater desalination unit, this means operating within a pH
range of 7.8–8.2, which is normal for seawater, i.e. without the alkali dosing that
is frequently applied to increase pH and so prevent calcium carbonate scaling. For
this so-called alkaline mode of operation up to a maximum pH of 9.0, though, an
efficient inhibitor must be dosed at a suitable rate to prevent scaling [57]. In this way,
boron rejection and also its temperature dependency can be appreciably improved.
However, this alkaline mode of operation of a seawater desalination membrane
desalination system is more or less limited to the application of TFC polyamide
membranes. For cellulose acetate membranes, their service lifetime may be mark-
edly reduced if operated at a pH of 7.5 and above (see Sect. 5.1.5.2.3, Fig. 5.22).
For further boron reduction in a second desalination stage equipped with brackish
water membranes, its feed pH may be increased even more up to a maximum of 10.5.
This means that the share of boron in the total boron content dominates and as a
consequence the total boron permeability is low with a correspondingly high boron
rejection.
In the second pass of an SWRO in particular, the reduction in the membrane’s
boron rejection rate due to ageing may be compensated by appropriately raising the
pH by dosing sodium hydroxide solution into the feed to the membrane system. This
is also possible in a seawater desalination membrane system, although to a lesser
degree as this is limited due to the increasing potential for scaling.
The curves of the two graphs in Figs. 5.25 and 5.26 are calculated on the basis of
the algorithms of Eqs. (5.94)–(5.95c). For this, it is assumed that the permeability of
the two components boric acid and borate will only be influenced by the temperature.
Investigations at brackish water membranes have shown, though, that for this
membrane type the permeability coefficients of the two components also themselves
depend on pH. However, the influence of the pH is very different depending on the
membrane manufacturer and membrane type [58]. In this case, the respective
membrane permeability coefficients for boric acid and borate may be multiplied
with correction factors that take account of their dependency on pH (Eq. 5.98).
Likewise identified was a dependency of the components’ reflection coefficient on
pH .

BB,EðStÞ ¼ F BðOHÞ3  f C,BBðOHÞ  BBðOHÞ3 ,EðStÞ þ F BðOHÞ4 


3 ,E ðStÞ

 f C,BBðOHÞ  BBðOHÞ4  ,EðStÞ ð5:98Þ


4 E ðStÞ

f C,BðOHÞ ,EðStÞ ¼ correction factor for B(OH)3 permeability dependence on pH or


3
salinity []
5.1 Reverse Osmosis Membranes and Fundamentals of Their Application 425

f C,BðOHÞ  EðStÞ ¼ correction factor for B(OH)4 permeability dependence on pH or


4
salinity []
The curves of the graphs in Figs. 5.25 and 5.26 therefore show the dependencies
of boron permeability on pH and temperature in part in idealized form. Especially
within the neutral and weakly alkaline pH range, as the investigations revealed,
boron permeability can increase again and it does not attain its lower values until a
higher pH range is reached. The reason for this membrane behaviour is primarily the
change in the surface charge (zeta potential) of the membrane separation layer
caused by the change in pH. As a consequence, a Donnan potential can build up
on the membrane surface with an increased influence on ionogenic components and
their transport through the separation membrane.
Boron rejection by RO membranes is a complex process with regard to the
physical and physico-chemical processes in the membrane separation layer and on
its surface as well as the dependency of the distribution ratio of the boron
components on pH. It is significantly influenced to a major part by specific membrane
properties, but also by the conditions at which the membranes are operated [59–
61]. Thus, for the selection of the most suitable membrane for a specific application,
the practical experience of the individual membrane manufacturers with their avail-
able membranes with increased boron retention in similar operating cases should
therefore be taken into account, or comparative tests should be carried out with
different membranes under the relevant operating conditions [62, 63].

5.1.5.2.5 Sequence of Membrane Element Design Calculations


Presented in Table 5.10 is an overview of the calculation sequence for determining
the operational parameters of production performance and product flux together with
the concentration of each component in the product and the overall product compo-
sition for specified design conditions and operational lifetime for a membrane
element from the standard test conditions of the element stipulated by the membrane
manufacturer. For this calculation sequence, the approach with the use of correction
factors for product flux and salt passage has been selected as it allows the use of the
standard test conditions specified by the membrane manufacturers for their elements
(see Sect. 5.1.5.2.3). The manufacturers normally do not publish the permeability
coefficients of their membrane elements for water transport AE(St) and salt transport
BiE(St) under standard conditions. However, as shown in Sect. 5.1.5.2.3, these can be
derived from the standard test conditions.
The algorithm to be used in each calculation step is named in the table together
with the relevant equations listed or derived in Sect. 5.1.5 and elsewhere in this book.
The calculation sequence shown may be generally applied for membrane
elements of differing types and designs. However, various equations allotted in
Table 5.10 to a certain calculation step, like for the temperature correction factors
for water and salt transport fTCW and fTCS, the concentration polarization factor βE,
the pressure loss across a membrane element ΔpFC,E as well as the correction factors
for the product and salt permeabilities f FP,E,τMOp and f FSi,E,τMOp for a stipulated
membrane operational lifetime, are specific to polyamide TFC spiral-wound
426 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.10 Sequence of calculation of design performance parameters for RO membrane


elements

(continued)
5.2 RO Membrane Process Design 427

Table 5.10 (continued)

Step Equation no. for


Settings to be available Parameter to be calculated
no. calculation
conditions and at operation time τMOp(d)

elements. Other element-specific coefficients for calculating these parameters apply


just to 800 elements. For other membrane element models or for membrane elements
with other membrane types, the calculation steps of the calculation procedure shown
must be assigned accordingly for these specific algorithms.

5.2 RO Membrane Process Design

5.2.1 RO Membrane Module and RO System Configuration

Critical for the recovery rate at which an RO desalination element can be operated is
its permissible degree of concentration polarization (see Sect. 5.1.5.2.2, Eqs. 5.53
and 5.55). Figure 5.27 shows for an 8" seawater desalination spiral-wound polyam-
ide element how, with increasing recovery YE of the element, the average concentra-
tion polarization factor βE and thus also the salt concentration at the membrane wall
increase, so impacting its salt rejection and production performance.
428 5 Reverse Osmosis Membrane System: Core Process of SWRO

Average concentration
polarisation factor βE of
element [-]
1.25

1.20
RO seawater polyamide spiral wound 8" element

1.15

1.10

1.05

1.00
0 5 10 15 20 25
Recovery YE of element [%]

Fig. 5.27 RO seawater 8" polyamide element—dependence of concentration polarization on


element recovery

As the graph shows, the limit for concentration polarization of 1.2 times the
concentration polarization factor βE as stipulated by the manufacturers of polyamide
TFC spiral-wound elements corresponds to a maximum RO element recovery of
around 21%. Additionally, commensurate with the quality of the feed to the RO
desalination system as well as how the raw water is extracted and its upstream
precleaning and pretreatment, an element’s recovery for seawater desalination is
limited by the membrane manufacturers to 10–15%, which corresponds to a βE range
of between around 1.10 and 1.15. This means that, when designing an RO desalina-
tion system, its membrane elements are to be dimensioned for a product recovery
within this range and in subsequent operation they may not be operated at a rate in
excess of this. For the brackish water elements for post-desalination in the second
pass, a recovery rate of up to 30% is possible (Table 5.12).
In order to attain an RO system recovery YS of 35–55% as required for commer-
cial seawater desalination and up to 90–95% for post-desalination in the second pass,
the membrane elements are connected in series on the concentrate side and assem-
bled to membrane modules for loading into pressure vessels. The discharge of
concentrate from an element then provides the feed to the element that is immedi-
ately downstream of it. The product header tubes of each element are likewise
connected in series to a central module product header tube in which the permeate
from each element is piped to the head ends of the membrane module (see Fig. 5.29
as well as Sect. 5.1.2, Figs. 5.5 and 5.12).
5.2 RO Membrane Process Design 429

For seawater desalination with polyamide spiral-wound elements, six to eight of


these elements are assembled to make up a membrane module within a pressure
vessel. Modules for brackish water desalination comprise up to six elements.
Owing to the higher recovery of cellulose acetate hollow-fibre modules, for these
systems usually membrane modules with two elements suffice to attain a maximum
system recovery of 50% for seawater desalination.
If the number of elements in a single membrane module is not enough to attain a
certain specified system recovery, there are various module configuration options for
achieving a recovery rate that exceeds the value possible with just one membrane
module (Fig. 5.28).

Concentrate recycle

Feed Concentrate
Membrane module

A Product

Product

Membrane module Membrane module

Membrane module Membrane module

Membrane module Membrane module

Concentrate
Feed

Product

Membrane
module
Membrane
Feed module
Membrane Membrane
module module
Membrane
module Concentrate
Membrane
module

Fig. 5.28 RO system configuration options


430 5 Reverse Osmosis Membrane System: Core Process of SWRO

By recirculating concentrate back to the feed supply to the membrane module, its
throughflow rate can be raised so that, even with an increasing system recovery YS,
the module’s recovery rate YM and that of its elements can be kept within its
elements’ permissible bandwidth YE (Fig. 5.28, Configuration A). The feed pump
then has to be dimensioned to provide the delivery head required to maintain the
supply rate for the module to attain the system’s specified permeate yield YS while
providing the pumping capacity needed for concentrate recirculation.
Such a configuration can be operated continuously as well as on a batch basis. For
the latter case, the concentrate from the membrane module is piped to a feed tank
whose content is concentrated by recirculating it through the module until the
specified recovery rate or concentration factor is attained at which point this hold
tank is refilled with raw water to provide the next batch. Operation with continuous
concentrate recirculation as well as with batch operation is to be found primarily in
pilot and trials plants. Also, desalination plants with a very low capacity are operated
in accordance with Configuration A.
Configuration B shows a two-stage plant in which the flow through the mem-
brane modules and their elements as needed for limiting concentration polarization is
attained by concentrate recirculation internally within the stages.
The feed pump for the membrane system is dimensioned to attain the desired
system recovery YS and maintain the operating pressure needed for the separation
process. The concentrate recirculation pumps for each stage must possess sufficient
pumping capacity for each stage to maintain the flow through the membrane
modules. Regarding their delivery head, they merely have to overcome the differen-
tial pressure of the respective stage. The specified plant production performance is
attained by connecting the number of membrane modules needed for this purpose in
parallel within the two stages. This staged Configuration B is encountered primarily
in micro- and ultrafiltration plants that are operated continuously in a so-called
crossflow mode. This configuration finds virtually no application in membrane
desalination plants.
With Configuration C, for limiting concentration polarization the flow through
the elements as needed for this purpose is matched to the fluctuating feed rate so as to
attain the desired production performance by distributing the required modules over
more than one stage, with each stage made up of differing numbers of membrane
modules. The number of modules and thus also the number of elements in each stage
is so matched to the reduction of product output due to the progressive rise in
concentration that successive stages will still exhibit similar recoveries. This is
done by an appropriate reduction of the membrane modules from the first stage of
the configuration to the subsequent second stage or, for brackish water desalination,
also to a third stage, should this be needed. Today, commercial plants for seawater
and brackish water desalination are almost entirely constructed with this membrane
module configuration in a pyramid-like or tapered arrangement. Such a staged
configuration of membrane modules is also known as an “array”, which, if carefully
designed, allows a completely continuous operation of the reverse osmosis system;
the compliance with the guidelines and limits set by the membrane manufacturers
over a wide range of temperatures of the water to be desalinated and also a
5.2 RO Membrane Process Design 431

pF,E1 FF,E1 pF,E2 FF,E2 pF,E3 FF,E3 pF,E4 FF,E4 pF,E5 FF,E5 pF,E6 FF,E6 pF,E7 FF,E7
YE1 cF,E1 YE2 cF,E2 YE3 cF,E3 YE4 cF,E4 YE5 cF,E5 YE6 cF,E6 YE7 cF,E7
FF,M πF,E1 πF,2 πF,3 πF,4 πF,5 πF,6 πF,7 FC,M
pF,M cF,M pC,M cC,M
πF,M πC,M
E1 E2 E3 E4 E5 E6 E7
Feed F Concentrate C

FPE3 cPE3 FPE4 cPE4 FPE7 cPE7 pP,M FP,M


FPE1 cPE1 FPE2 cPE2 FPE5 cPE5 FPE6 cPE6
YM cP,M
πP,M

Product P

Fig. 5.29 RO membrane module with series-connected elements

Stage 1
Array

M 1.1 Stage 2
1 2 3 4 5 6 7 Concentrate CA
FF,A,St1 FF,St2 M 1.2
pF,A;St1 cF,A,St1 FC,A
cF,St2
M 2.1 1 2 3 4 5 6 7 pC,A cC,A
πF,A,St1 πF,St2
πC,A
1 2 3 4 5 6 7 M 2.2
Feed FA pF,St2
M 3.1 1 2 3 4 5 6 7

1 2 3 4 5 6 7
YSt1 FP,St1 YSt2 FP,St2
Product PSt1 cP,St1 Product PSt2 cP,St2
πP,St1 πP,St2 Product PA

YA FP,A
pP,A cP,A
πP,A

Fig. 5.30 RO system staging

corresponding adaptation of the operating conditions to fluctuating salinity of the


raw water.
The arrangement of several membrane elements to make up a membrane module
and the associated mass balance symbols are shown in Fig. 5.29. The module shown
contains seven elements as is frequently encountered in seawater desalination plants.
To attain the specified capacity, these modules are connected in parallel in a
single- or multi-stage array. For seawater desalination, predominantly a single-stage
array is employed with a two-stage array needed only in exceptional cases. Such a
two-stage configuration is shown in Fig. 5.30, again with the associated mass
balance symbols.

5.2.2 Design Phases of an RO System

The design of an RO membrane system and calculation of the performance data of its
membrane elements and modules as well as of the system array as a whole proceeds
in four phases:
432 5 Reverse Osmosis Membrane System: Core Process of SWRO

• Phase 1—Definition of design conditions and basic design parameters


• Phase 2—Calculation of the framework parameters of the membrane system and
the start-up values as inputs for optimization and detail calculation of the baseline
system design
• Phase 3—Optimization of the RO system configuration, detail calculation of the
performance of the components, and of the composition of the product generated
by membrane elements, modules, stages, and arrays
• Phase 4—Extension of the basic design to cover the full ranges of TDS, compo-
sition, and temperature of the feed.

5.2.2.1 Phase 1: Definition of Design Conditions and Basic Design


Parameters
In this phase, the performance data for the design of the membrane system are laid
down; the operating conditions and the basic design parameters for its dimensioning
are determined; membrane elements and their technical data are selected; and
technical parameters for the module configuration are defined (Table 5.11).
From the bandwidths of composition and temperature of the feed to the mem-
brane system, design points are defined for both these parameters for preparing a
basic design. Usually, the mean of the minimum and maximum values is selected for
this purpose. Because the operating conditions and product composition of an RO
desalination system depend to a large degree on feed temperature and TDS, follow-
ing preparation and optimization of the basic design in Phase 3, in Phase 4 the
calculation is extended to encompass the whole range of feed conditions and,
depending on the outcome of this, the design is fine-tuned as necessary regarding
plant configuration and operating conditions. In doing so, the system performance
parameters for the respective feed conditions and the related operating cases are
documented.
Alongside the performance parameters that have to be taken into account for the
design, i.e. product flow and quality, together with the feed and operating conditions
that govern plant output, the average specific product flux JP,SØ,S is critical for
dimensioning and configuring the membrane system. This is selected in line with
the extraction and pretreatment conditions for the seawater to be desalinated from the
bandwidths given in Table 5.12 for this parameter and is calculated using Eq. (5.99)
from the system product flow FP,S and the installed system membrane area SS, which
is equal to the number of membrane elements present in the system N ES times their
specific membrane area SE.

F P,S  1000 F P,S  1000


J P,S∅,S ¼ ¼ : ð5:99Þ
SS N E S  SE

JP, S ∅ , S ¼ average specific product flux of RO system ¼ JP, S ∅ , M for


membrane module [l/m2h] ¼JP, S ∅ , St for stage [l/m2h] ¼ JP, S ∅ , A for array
[l/m2h]
SS ¼ membrane area of system [m2]
SE ¼ membrane area of element [m2]
N ES ¼ total number of elements of system []
5.2 RO Membrane Process Design 433

Table 5.11 Design phase 1—definition of design conditions and setting of basic design
parameters
Parameter Symbol Unit
Performance parameter of system
Production performance; product flow CP,S; FP,S m3/day, m3/h
Product recovery factor YS –
Product component concentration targets of system ciPS g/l, kg/m3
Product salt content or TDS cPS
Operation design conditions
Feed analysis, components concentration g/l, kg/m3
• Design point for baseline design ciFS(d )
• Minimum ciFS(min)
• Maximum ciFS(max)
• Average ciFS(aver)
Feed temperature

• Design point for baseline design tF(d ) C
• Minimum tF(min)
• Maximum tF(max)
• Average tF(aver)
Feed quality—SDI or MFI SDIF; MFIF –
Performance parameter and make of elements (see Tables 5.37, 5.38, and 5.39)
Category Normal; high rejection; high flux; low
fouling; etc.
Type Manufacturer-specific
Size Depending on manufacturer and system
capacity
Standard production performance and flow CP,E(St); FP,E(St) m3/day, m3/h
Membrane surface area SE m2
Element spacer thickness hSp mil; mm
Membrane configuration design parameters
Average specific product flux JP,SØ,S l/m2h
Concentration polarization factor of element, max. βE,max –
Recovery of element, max YE,max –
Membrane lifetime τMOp, E Years
Membrane module design parameters
Elements per module NE,M –

By specifying this parameter and from the desired product flow FP,S of the
system, it is possible to estimate the number N ES of membrane elements that will
presumably be required.
But the number of membrane elements depends also on what element type is to be
installed with its product flow under standard conditions together with its membrane
area. Element selection is influenced by the composition and quality of the feed to
the membrane system and the specified product quality. Particularly significant for
434

Table 5.12 Design guidelines and limits for TFC polyamide spiral-wound membranes
Manner of extraction, pretreatment, and product type
5

Sea open—
conventional Sea open—advanced Sea well or sea RO
Parameter Symbol Unit Condition pretreatment conventional pretreatment open—MF/UF permeate
Average product flux of JP,SØ,S l/m2h Range 11–17 14–17 13–22 30–43
system max. 17 17 22 45
Lead element product flux JP,E1 l/m2h max. 28–34 34 35–42 48–56
Recovery of element YE % max. 10–13 14 12–15 30
Feed flow to element FFE m3/h max. 13–17 13.5–15 14–17 15–17
Brine flow of element FCE m3/h min. 2.7–3.6 3.2 2.7–3.6 2.3–2.7
Concentration polarization βE – max. 1.2 – 1.2 1.3–1.5
factor of element
Silt density index SDI %/min Range <3 <3 1–2 <1
max. 4–5 3 3 1
Reverse Osmosis Membrane System: Core Process of SWRO
5.2 RO Membrane Process Design 435

this is the fouling potential of the raw water (SDI/MFI) subsequent to its
pretreatment, but also its residual organic load and its still remaining potential for
biological fouling (see Sect. 4.2.1.1).
For identical element dimensions, spiral-wound modules with differing
characteristics are available, like:

• high salt and boron rejection


• high product flow and relatively low salt rejection
• enhanced tolerance to fouling due to greater element spacer thickness, but in part
at the cost of a lower product flow, coupled with special structuring of the
element’s membrane surface to reduce the adhesion of particulate deposits.

(see Sect. 5.1.5.2.3, Tables 5.4, 5.37, 5.38, and 5.39). Membrane elements are
then selected to match the quality of the feed to the membrane system as well as the
system’s target product flow and quality, but also under the aspect of optimizing the
desalination plant’s energy consumption.
A further critical parameter for calculating membrane performance is their ageing
behaviour, i.e. the service lifetime τMOp, E specified for the element. This applies for
both product flow and for the membrane’s salt rejection, i.e. the product TDS and the
concentration of its constituents.
As well as the need to limit the concentration polarization factor βE, other
standards and limits stipulated by the membrane manufacturers must be complied
with when dimensioning the series-connected and in concentrate staging arranged
elements. Such parameters relevant for polyamide spiral-wound modules are listed
in Table 5.12.
To a large part, these parameters will be specified so as to be absolutely sure that
there will be no elevated concentration polarization in the elements, but also to
ensure uniform throughflow. A further criterion is that the flow rate through the
module will not exceed a certain value and thus the permissible differential pressure
across the element and the membrane module will not be exceeded (see Sect.
5.1.3.1.1). The various guide and limit values to be applied for dimensioning the
membrane elements are assigned to meet the extraction and pretreatment
requirements of a seawater desalination system under the aspect that, in consider-
ation of differing feed conditions and the resulting variation in fouling potential,
deposits of swept-in colloidal and particulate foulants on the element membranes are
to be minimized (see Sects. 5.3 and 5.3.2.1). Still other target values have to be taken
into consideration when dimensioning membrane elements for performance optimi-
zation during Phase 3 of the design process. These are selected from the prescribed
ranges of values compiled in Table 5.12 to match the manner of seawater extraction
and the planned pretreatment systems.
436 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.2.2.2 Phase 2: Calculation of the Framework Parameters


of the Membrane System and of Approximate Values as Inputs
for Optimization and Detailed Calculation of the Baseline
System Design
During this phase, an initial estimate is made of the number N ES of membrane
elements needed for the stipulated system product flow FP,S and, on the basis of the
specified system recovery YS, the anticipated configuration of these elements in
membrane modules is determined as well as how these modules are arranged in
just one or more stages within an array depending on the design conditions.
The system recovery YS that is specified for the design of the RO system not only
determines the configuration of its membrane elements and modules, but also the
scaling potential of the concentrate of the desalination plant resulting from the rise in
concentration of sparingly soluble seawater constituents, i.e. scalants, during the
desalination process. But measures for scaling control can affect the degree of
rejection for specific seawater constituents and thus the ionogenic composition of
the generated product water, depending on whether both acid and antiscalant are
dosed to suppress scaling, so reducing the seawater’s pH, or whether scaling is
controlled without changing pH by just dosing antiscalant (see Sect. 5.1.5.2.4).
For this reason, calculations for scaling control (see Sects. 5.3.4 and 5.3.4.2) and
any changes in the ionogenic composition of the feed to the RO system resulting
from this are also part of this design phase.

5.2.2.2.1 Estimates of Product Recovery, Staging, and Feed Pressure


Estimates of Product Recovery and Staging
If separation units are connected in series within modules for concentrate staging as
shown in Fig. 5.29 or if more than one of these modules containing separation units
are connected in an array staging configuration in accordance with Fig. 5.30, the
recovery rate YS of the module or array system may be calculated from the recovery
rate YUi of each membrane unit and the number NU of the interconnected units by
applying Eq. (5.100).

YS ¼ 1
" #
Y
i¼n
 ð1  Y U 1 Þ  ð1  Y U 2 Þ . . . ::  ð1  Y U n Þ ¼ 1  ð1  Y U i Þ
i¼1

ð5:100Þ

YUi ¼ recovery rate of membrane unit n []


YS ¼ recovery factor of system []
n ¼ number of units []
Assuming the same recovery rate YU for all units Ui or taking the mean of their
recovery rates Y U , with Eq. (5.101), a geometrical series can be developed on the
basis of the recovery rate of the unit Ui.
5.2 RO Membrane Process Design 437

 n1
F P,U i ¼ F P,U 1  1  Y U : ð5:101Þ

FP, Ui ¼ product flow of unit i in series-connected element train [m3/h]


FP, U1¼ product flow of first unit in series-connected element train [m3/h]
Y U = average recovery rate of all units []
By taking the sum of this geometrical series together with Eq. (5.101a) for the
system recovery YS, Eq. (5.101b) results and from this Eqs. (5.101c) and (5.101d)
are derived.

F P,S
YS ¼ : ð5:101aÞ
F F,S

FP,S ¼ product flow of system [m3/h]


FF,S ¼ feed flow of system [m3/h]
 N
YS ¼ 1  1  YU U : ð5:101bÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Y U ¼ 1  ð1  Y S Þ =N U = 1  N U 1  Y S : ð5:101cÞ
1

log ð1  Y S Þ
NU ¼  : ð5:101dÞ
log 1  Y U

Y U ¼ mean recovery factor of all units Un in the membrane system []


NU ¼ number of units connected in series []
With Eq. (5.101b), the anticipated product recovery YS of the system can be
calculated from the average recovery of the separating elements Y U and their number
NU. Eq. (5.101c) accordingly yields the mean product recovery rate Y U
corresponding to the system recovery YS and the number of units NU. Equation
(5.101d) permits calculation of the number of separation units NU that are needed to
attain a defined system recovery YS at a given mean recovery rate of the units Y U
[64].
The interdependency of the three parameters system recovery rate YS, mean
product recovery Y U of the separation units, and the number NU of series-connected
units is shown graphically in Fig. 5.31. The range of up to 10% of the mean recovery
rate Y U and up to a number NU of 6 to 8 of series-connected units corresponds to the
situation for spiral-wound elements for seawater desalination.
Equation (5.101b) shows that with Y U values of 8% to 10% and by equipping a
module with seven elements, system recovery rates of 42% and 50% are obtained.
With an NU of eight elements per membrane module, under these conditions a
system recovery rate YS of between 50% and 55% is calculated. To obtain a higher
system recovery with the same number of seven or eight elements per membrane
module, a staged configuration is necessary. If a system recovery of 60% is to be
obtained with a two-stage array, then each stage would have to have a recovery rate
438 5 Reverse Osmosis Membrane System: Core Process of SWRO

Recovery rate YS of
module, stage or
system [%]
110

100 Number of
serial units
90 NU
8
80 7
6
70
5
60 4

50
3

40
2
30

20

10

0
0 10 20 30 40 50 60 70
Average recovery rate of serial unit YU ( element, module or stage) [%]

Fig. 5.31 Dependence of system recovery rate YS on the mean unit recovery rate Y U and number of
units NU

of approximately 36%. With CTA hollow-fibre modules that are normally equipped
with two membrane units per module, a mean recovery rate per element of 25 to 30%
is needed to obtain a system recovery rate of 45 to 50%. To attain system recovery
rates of 90% and more, that is values for which the second pass of a seawater
desalination plant is usually dimensioned, arrays with up to three stages are needed
with the stages themselves each having to exhibit recovery rates from 55% to over
60%.
By calculating the quotient qSt of the number NU of units to be connected in series
to obtain a system recovery YS and the number NE,M of elements selected for each
membrane module, it is possible to estimate whether a defined system recovery YS
can be obtained with just a single stage or if more than one stage would be needed
(Eq. 5.102).

NU
qSt ¼ : ð5:102Þ
N E,M

qSt ¼ staging quotient []


NE,M ¼ number of elements per module []
If this staging coefficient qSt works out to greater than unity, then either the
number of elements per module NE,M will have to be increased or, if this is not
possible, a two-stage module configuration has to be selected. Should the value of
the staging coefficient qSt be in the vicinity of two or more, a three-stage configura-
tion of the RO system will be needed.
5.2 RO Membrane Process Design 439

Table 5.13 Staging of seawater and brackish water RO systemsa


Number of stages NSt with
System Number of series-connected 6-element 7-element 8-element
recovery YS elements NU module module module
% [] [] [] []
Seawater desalination
35–40 6 1 1 –
45 7–12 2 1 1
50 8–12 2 2 1
55–60 12–14 2 2 –
Brackish water desalination
40–60 6 1
70–80 12 2
85–90 18 3
a
Source: Dow Water & Process Solutions—FILMTEC Reverse Osmosis Membranes—Technical
Manual

However, then the algorithms as described, which are based solely on mass
balance considerations for series-connected elements together with their mean
recovery rates, can only yield indicative values for the configuration of membrane
elements in modules and how these modules are to be arranged in stages and arrays.
Nevertheless, it is apparent that the results so obtained are in good agreement with
figures for configuring desalination plants as published by membrane manufacturers
based on their experience and the results of analyses using their design software (see
Table 5.13).

Estimate of Feed Pressure


In order for permeate to flow through an RO membrane, it has to be subjected to a
differential pressure that exceeds the prevailing osmotic pressure difference. This
basic principle of reverse osmosis is also termed the thermodynamic limit or
thermodynamic restriction. This relationship defines the limit for the application of
reverse osmosis for desalination of seawater of specific TDS by reason of the
potential concentration increase and the system recovery rate coupled with the
required feed pressure, i.e. the differential pressure.
For an RO system with membrane elements connected in series, because of the
thermodynamic limit the feed pressure pF,N ES to the final element in the system has to
be greater than the difference Δπ S, C, m, P of the osmotic pressure at its membrane
wall and that of the RO system product (Eq. 5.103).

pF,S  pP,S  ΔpFC,S ¼ pF,N E  Δπ S,C,m,P  π S,C,m,  π P,S


S

 π F,S  CFS  βS  π P,S ð5:103Þ


440 5 Reverse Osmosis Membrane System: Core Process of SWRO

pF,S ¼ feed pressure of system ¼ pF,M for membrane module ¼ pF,A for RO array
[bar]
pP,S ¼ product pressure of system [bar]
pF,N E ¼ feed pressure at last element N ES of system [bar]
S
ΔpF, C; S ¼ feed to brine pressure loss of system ¼ ΔpF, C; M ¼ for membrane
module ¼ ΔpF, C; A for RO array [bar]
Δπ S, C, m, P ¼ osmotic pressure difference of system between concentrate at
membrane wall at last unit n N ES and product ¼ Δπ M, C, m, P for membrane module ¼
Δπ A, C, m, P for RO array [bar]
Δπ S, C, m = osmotic pressure of concentrate of system at membrane wall at last
unit n N ES ¼ Δπ M, C, m for membrane module ¼ Δπ A, C, m for RO array [bar]
π P,S = osmotic pressure of system product ¼ π P,M for membrane module ¼ π P,A
for RO array [bar]
π F, S ¼ osmotic pressure of system feed ¼ π F, M for membrane module ¼ π F, St for
RO stage [bar]
βS ¼ average system concentration polarization factor ¼ βM for membrane
module ¼ βA for RO array []
CFS ¼ system concentration factor ¼ CFM for membrane module ¼ CFA for RO
array []
If the differential pressure at the system’s final element is equated to the differ-
ence of the osmotic pressure there prevailing, with Eqs. (5.103) and (5.103a) and
under the simplifying assumption that for seawater desalination the osmotic pressure
in the product π P,S is very small in comparison with that of the concentrate π S, C, m,
there results a relationship for the maximum or limiting system recovery rate YS,max
(Eq. 5.103b).
 
1  Y S  1  RS
CFS ¼ : ð5:103aÞ
1  YS
for π P,S π S, C, m

π F,S  βS  RS
Y S, max ¼ 1  ð5:103bÞ
pF,S  ΔpF,C; S  pP,S

YS, max ¼ maximum system recovery rate from thermodynamic limit []
RS ¼ average system salt rejection rate ¼ RM for membrane module ¼ RA for RO
array []
For seawater with a defined TDS cF,S and a consequent osmotic pressure π F,S in
the feed to the RO system, for a predefined feed pressure pF,S the system recovery
rate YS has to be less than this limit value YS,max.
Similarly, a relationship may be derived from Eq. (5.103b) with which a limiting
minimum feed pressure pF,S,min to the RO system can be determined if, for a specific
seawater composition, the system recovery rate YS is predefined. The feed pressure
pF,S must then be higher than the limiting minimum pressure pF,S,min (Eq. 5.103c).
5.2 RO Membrane Process Design 441

π F,S  βS  RS
pF,S, min ¼ þ ΔpF,C; S þ pP,S : ð5:103cÞ
1  YS
pF,S,min ¼ min. system feed pressure from thermodynamic limit ¼ pF,M,min for
membrane module ¼ pF,A,min for RO array [bar]
The relationship between feed TDS and temperature of seawater as well as the
corresponding thermodynamically limiting parameters, the maximum system recov-
ery rate YS,max and the minimum feed pressure pF,min are graphed in Fig. 5.32. These
curves are calculated for a mean salt rejection RS of 99.6%, which corresponds to a
seawater desalination membrane with a standard rejection rate of 99.8% for NaCl
and a mean concentration polarization factor βS of 1.1 for the RO system. Not
considered are the backpressure pP,S and the pressure loss ΔpF, C; S in the system.
If, for example, for seawater with a TDS of 30 g/kg and a temperature of 20  C, a
system recovery of 50% (YS ¼ 0.5) is aimed for, the limiting feed pressure pF,min
amounts to 48 bar. In contrast, for a system recovery rate of 40% (YS ¼ 0.4) for a
seawater TDS of 50 g/kg and a temperature of 30  C, the limiting feed pressure will
have already increased to 70 bar. Because at 30  C the permissible maximum
operating pressure of polyamide TFC seawater RO membranes is limited by the
membrane manufacturers to 75 bar (Fig. 5.13) and, to match the membrane
characteristics, the limiting feed pressure pF,min has still to be raised to a practical
feed pressure pF,S, at the given operating conditions it will presumably not be

Limiting system
recovery factor YS,max [-]
0.9
Feed temperature [°C ] Feed salinity [g/kg]
10
0.8 20
20
30
0.7 40 30

0.6 40

0.5
50

0.4

0.3

0.2

0.1
10 20 30 40 50 60 70 80 90
Minimum feed pressure pFmin [bar]

Fig. 5.32 Thermodynamic limit of seawater RO—maximum system recovery and necessary
minimum feed pressure for various feed salinities and temperatures
442 5 Reverse Osmosis Membrane System: Core Process of SWRO

possible to attain this system recovery rate without exceeding the membrane
element’s permissible operating pressure.
The amount by which the system recovery YS,max will have to fall short of its
specified value, or alternatively, the feed pressure pF,min will have to be exceeded to
attain such a rate depends, for the selected membrane type, on its specific water
permeability and the influence of temperature and membrane ageing on this param-
eter as well as the concentration polarization and salt rejection within the RO system.
Experience shows that for an RO seawater desalination system, depending on the
selected membrane type and specific product flux JP, S ∅ , S, its feed pressure will be
around 1.05–1.15 times the osmotic pressure at the membrane wall of the final
element. An estimate pF, S, estim derived from pF, S, min together with its empirical
factor f pFestim in accordance with Eq. (5.103d) may be taken as starting value for the
subsequent detailed optimization calculation for the RO system. Half of the maxi-
mum permissible pressure loss of a membrane module ΔpF, C, Emax as quoted by the
membrane manufacturer (see Table 5.2) multiplied by the number of elements that
are connected in series N ES should be taken as starting value for the system pressure
loss ΔpF, C; S in Eqs. (5.103c) and (5.103e).

pF,S,estim ffi pF,S, min  f pFestim : ð5:103dÞ

f pFestim ¼ 1:05  1:15

ΔpF,C,E max
ΔpF,C; S,estim ¼  N ES ð5:103eÞ
2
pF, S, estim ¼ estimate of system feed pressure as starting value for optimization
calculation [bar]
ΔpF, C; S, estim ¼ estimate of system pressure loss as starting value for optimiza-
tion calculation [bar]
ΔpF, C, Emax ¼ maximum pressure loss quoted by membrane element manufac-
turer [bar]
N ES ¼ number of elements in system ¼ N EA of array []

5.2.2.2.2 Estimates of Number of Membrane Elements and Modules


with Module Stage Distribution
The number of elements N ES that the RO system will have to be equipped with to
attain the stipulated product capacity CP,S or product flow FP,S can be estimated from
the system’s selected specific product flux JP, S ∅ , S and the membrane area SE of the
chosen membrane element type using Eq. (5.104). An approximation for the total
number of membrane modules N M S needed for the system can be calculated from the
number of membrane elements per module N ES and the selected number of elements
for installation in the modules NE,M (Eq. 5.105).
5.2 RO Membrane Process Design 443

C P,S  1, 000 F  1, 000


N ES ¼ ¼ P,S : ð5:104Þ
24  J P,S∅,S  SE J P,S∅,S  SE
N ES
N MS ¼ : ð5:105Þ
N E,M

CP,S ¼ product capacity of RO system [m3/day]


FP,S ¼ product flow of RO system [m3/h]
N M S ¼ total number of modules of system/array (pressure vessels) []
NE,M ¼ number of membrane elements per module []
With recovery rates for RO seawater desalination systems YS of 35–50%,
depending on the seawater TDS, normally a multi-stage configuration of the mem-
brane modules is not needed and all of the system’s membrane modules N M S are
arranged in just one array, i.e. the number of stages NSt is 1 and the number of
modules in this single-stage array N M,St1 works out to NM, A (or N M S ).
Within an array configuration comprising several stages, the distribution of the
number of membrane modules N M S in the system is fixed by a staging ratio RSt. This
is the ratio of the number of modules in the respective stage of the array to the
number in the subsequent stage (Eq. 5.106) [65].

N M,Sti
RSt ¼ : ð5:106Þ
N M,Stiþ1

RSt ¼ staging ratio []


N M,Sti ¼ number of modules in stage Sti []
Assuming that the system recovery YSt of all stages of an array is the same, the
staging ratio RSt is calculated from the system recovery YS and the number of stages
NSt in the array, as shown in Eq. (5.107).
 1=N
1 St
RSt ¼ ð5:107Þ
1  YS

NSt ¼ number of stages of array []


For a multi-stage array, the number of modules in the first stage N M,St1 is
given by:

N MS
for N st¼ 2 N M,St1 ¼ ð5:107aÞ
1 þ RSt 1
N MS
for N st¼ 3 N M,St1 ¼ ð5:107bÞ
1 þ RSt 1 þ RSt 2
and for the number of modules in subsequent stages:
444 5 Reverse Osmosis Membrane System: Core Process of SWRO

N M,Sti1
N M,Sti ¼ ð5:107cÞ
RSt

5.2.2.3 Phase 3: Optimization of the RO System Configuration, Detail


Calculation of the Performance of the Components
and the Composition of the Product Generated by Membrane
Elements, Modules, Stages, and Arrays
The performance data defined in Phase 1 plus the there stipulated operating
conditions and basic design parameters for the membrane system as well as the
technical details and dimensioning limits of the selected membrane elements
together with the subsequently determined data in Phase 2 for the numbers of
membrane elements and of membrane modules, their configuration and staging as
well as the estimated feed pressure serve in this Phase 3 as starting values for the
most iterative calculations for detail optimization of the membrane desalination
system, its component configuration, and its performance data.

5.2.2.3.1 Series-Connected Element Flow, Pressure, and Concentration


Conditions
If as is shown in Fig. 5.29 membrane elements are series-connected in concentrate
staging to modules or if more than one of these element-containing modules are
interconnected in an array staging configuration like in Fig. 5.30, from the sum of the
product flows FP, Ei of the serially connected units, the product flow FP,A may be
calculated for the resulting array (Eq. 5.108).

i¼N EA
X
F P,A ¼ F P,E i ð5:108Þ
i¼1

FP,A ¼ product flow of array [m3/h]


FP, Ei ¼ product flow of membrane element Ei [m3/h]
The performance of a membrane element Ei with regard to product flow, recov-
ery, and composition depends on the operating conditions that become established at
its location within the array’s series-connected chain of elements. In order to
determine the product flow and product composition of a membrane module or of
an array, the performance data for each of its series-connected elements have to be
calculated in steps in the sequence in which they are installed and for the operating
conditions prevailing at their location. The performance figures for the system as a
whole are then obtained from the results of each membrane element.
The required feed flow FF,A to the array is then calculated from its product flow
FP,A, the system recovery YS (Eq. 5.109), and the concentrate flow of the array FC,A
with Eq. (5.110).
5.2 RO Membrane Process Design 445

F P,A
F F,A ¼ : ð5:109Þ
YS

1
F C,A ¼ F P,A  1 : ð5:110Þ
YS

FF,A ¼ feed flow to RO membrane array [m3/h]


YS ¼ recovery factor of RO membrane system/array []
FC,A ¼ concentrate flow of array
The feed flow F F,M , Sti to one of the membrane modules connected in parallel in
the respective array stages is obtained as the quotient of the feed flow to this stage
F F,Sti and the number of membrane modules N M,St1 installed there. For a single-
stage array as is usual for seawater desalination, the feed flow to a membrane module
is calculated in accordance with Eq. (5.111). For a multi-stage array, Eq. (5.111a)
applies.

F F,A
F F,M , St1 ¼ ð5:111Þ
N M,St1

F F,Stiþ1 F F,Sti  ð1  Y Sti Þ


F F,M , Stiþ1 ¼ ¼ ð5:111aÞ
N M,Stiþ1 N M,Stiþ1

F M , St1 ¼ feed flow to module i of stage i of array [m3/h]


Y Sti ¼ recovery factor of stage i
The feed flow FF, Ei to a series-connected element ni is equal to the feed flow
F M , Sti to the membrane module minus the sum of all product flows FP, Ei of the
upstream membrane elements ni1 (Eq. 5.112). Because of the elements’ series-
connected configuration, the concentrate discharge FC, Ei  1 from an element is
equal to the feed flow F FEi to the unit downstream of it (Eq. 5.112a). This means that
the feed flow of an element also depends on the product flow FP, Ei  1 and the
product recovery rate Y Ei1 of its preceding element (Eqs. 5.112b and 5.112c). The
ratio of the product flows of two consecutive units FP, Ei  1und FP, Ei is fixed in turn
by their product recoveries Y Ei1 und Y Ei in accordance with Eq. (5.112c).

X
i¼n i 1
F P,M X
i¼n i 1

F F,E i ¼ F F,M , Sti  F P,E i ¼  F P,E i ð5:112Þ


i¼1
Y Sti i¼1

F C,E i1 ¼ F FEi1  ð1  Y Ei1 Þ ¼ F FEi ð5:112aÞ



1
F F,E i ¼ F P,E i1  1 ð5:112bÞ
Y Ei1

F P,E i 1
¼ Y Ei  1 ð5:112cÞ
F P,E i1 Y Ei1
446 5 Reverse Osmosis Membrane System: Core Process of SWRO

FF, Ei ¼ feed flow to unit i [m3/h]


FP, Ei ¼ product flow of element i [m3/h]
Y Ei ¼ recovery factor of element i []
ni ¼ consecutive number of series-connected element []
Decisive for the product flow of a series-connected element is the net driving
pressure NDPEi acting on it, which is calculated from the feed pressure gradient at
the membrane element and the difference of the osmotic pressure at the membrane
wall (Eq. 5.113).

ΔpFC,Ei  
NDPEi ¼ pF,Ei  pP,A   π F,Ei  CFF,C,Ei  βEi  π P,A ð5:113Þ
2
pF,Ei ¼ feed pressure to series-connected element i [bar]
NDPEi ¼ net driving pressure at element i [bar]
ΔpFC,Ei ¼ pressure differential of feed to concentrate of element i [bar]
π F,Ei ¼ osmotic pressure of feed to element i [bar]
CFF,C,Ei ¼ average concentration factor feed to brine of element i []
βEi ¼ concentration polarization factor at element i []
πP,A ¼ osmotic pressure of product of element i [bar]
The feed pressure pF,Ei to the series-connected element Ei is calculated from the
feed pressure pF,A to the membrane module minus the pressure losses ΔpF,C,Ei in all
units ni1 upstream of the element ni1 together with the loss in feed pressure ΔpF, V
of the membrane pressure vessel.

X
i¼ni 1

pF,Ei ¼ pF,A  ΔpF,V  ΔpF,C,Ei ð5:114Þ


i¼1

pF,A ¼ feed pressure to array [bar]


ΔpF, V ¼ pressure loss at module vessel inlet [bar]
ΔpF,C,Ei ¼ pressure differential of feed to concentrate of element i [bar]
But in turn, the total pressure loss of an elementΔpF,C,Ei is made up of the
pressure loss of the element in its clean condition ΔpF,C,Ei,clean as calculated, for
instance, by Eq. (5.60a) for an 800 spiral-wound element plus an operative pressure
lossΔpFC,Ei,operation resulting from deposits in the flow cross-section whose value
depends on the frequency and thoroughness of membrane cleaning (Eq. 5.114a).
The total pressure loss may not exceed the maximum value for the membrane
element as prescribed by the manufacturer.

ΔpF,C,Ei ¼ ΔpFC,Ei,clean þ ΔpFC,Ei,operation ð5:114aÞ

ΔpF,C,Ei  ΔpF,C,E max

If the membrane element concerned is in the second stage of an array, its feed
pressure pF,Ei,St may be calculated as shown in Eqs. (5.114b)–(5.114f). In this case,
2
5.2 RO Membrane Process Design 447

the feed pressure to the array has to have subtracted from it the entire pressure loss of
the first stage plus the loss in the feed and discharge lines of the first stage membrane
pressure vessel together with that of the piping connecting it to the second stage.

X
i¼n i 1

pF,Ei,St2 ¼ pF,A  ΔpF,C,St1  ΔpF,V,St2  ΔpF,C,Ei, St2 ð5:114bÞ


i¼1

i¼N ESt
X1
ΔpF,C,St1 ¼ ΔpF,C,Ei, St1 þ ΔpF,VþP,St1 ð5:114cÞ
i¼1

i¼N ESt
X1
ΔpF,C,Ei, St1 ¼ ΔpF,C,M,St1  ΔpF,C,M max ð5:114dÞ
i¼1

ΔpF,C,Ei,St1 ¼ ΔpFC,Ei,clean,St1 þ ΔpFC,Ei,operation,St1 ð5:114eÞ

ΔpF,C,Ei, St2 ¼ ΔpFC,Ei,clean,St þ ΔpFC,Ei,operation,St ð5:114fÞ


2 2

ΔpFC,Ei,clean ¼ pressure differential of feed to concentrate of clean element i [bar]


ΔpFC,Ei,operation ¼ pressure differential of feed to concentrate of operated element
i [bar]
ΔpF, C, Emax ¼ maximum pressure differential of element [bar]
ΔpF, C, Mmax ¼ maximum pressure differential of module [bar]
ΔpF,C,Ei,Sti ¼ pressure differential of feed to concentrate of element i in stage
i [bar]
∑ΔpF, C, V + P ¼ sum of pressure differential of module vessels and piping [bar]
ΔpF,C,Sti ¼ pressure differential of stage i [bar]
The increase in concentration Δci,Ei of each solvent component due to the
separation process within the membrane element Ei depends on its product
recoveryY Ei and its rejection factor Ri,Ei for the respective component
i (Eq. 5.115). The concentration of the component ci, C, A in the concentrate
discharge of the array is given by Eq. (5.115b) from its feed concentration ci,F,Ei
plus the total increases in concentration within the membrane elements within the
array.

Y Ei  Ri,Ei Y E  Ri,Ei
Δci,Ei ¼ ci,F,Ei  ¼ ci,C,Ei1  i : ð5:115Þ
1  Y Ei 1  Y Ei

X
i¼ni
ci,C,Ei ¼ ci,F,A þ Δci,Ei : ð5:115aÞ
i¼1
448 5 Reverse Osmosis Membrane System: Core Process of SWRO

i¼NEA
X
ci,C,A ¼ ci,F,A þ Δci,Ei : ð5:115bÞ
i¼1

Δci,Ei ¼ concentration gradient of component i in element i [kg/m3, g/l]


ci,F,Ei ¼ concentration of component i in feed to element i [kg/m3, g/l]
ci,C,Ei ¼ concentration of component i in concentrate of element i [kg/m3, g/l]
ci, F, A ¼ concentration of component i in feed to array [kg/m3, g/l]
ci, C, A ¼ concentration of component i in array concentrate [kg/m3, g/l]
Ri,Ei ¼ rejection factor of component i in unit i []
Similarly, the increase of osmotic pressure Δπ Ei in each element of the membrane
module and the osmotic pressure in the concentrate discharge of an element i and of
the array π C, A may be calculated (Eqs. 5.116, 5.116a, and 5.116b).

YEi  Ri,Ei YE  Ri,Ei


Δπ Ei ¼ π F,Ei  ¼ π C,Ei1  i : ð5:116Þ
1  Y Ei 1  Y Ei

X
i¼ni
π C,Ei ¼ π F,A þ Δπ Ei : ð5:116aÞ
i¼1

i¼N EA
X
π C,A ¼ π F,A þ Δπ Ei : ð5:116bÞ
i¼1

π C,Ei ¼ osmotic pressure in concentrate of element i [bar]


π C, A ¼ osmotic pressure in concentrate of array [bar]
π F, A ¼ feed osmotic pressure of array [bar]
Δπ Ei ¼ osmotic pressure gradient in element i [bar]
π F,Ei ¼ feed osmotic pressure of element i [bar]
Following calculation of the net driving pressure NDPEi of the series-connected
element, its product flux J P,Ei and product flow F P,Ei may be determined as shown in
Eqs. (5.117) and (5.117a).

NDPEi
J P,Ei ¼ J P,EðStÞ,V  f TCWðdÞ  f FP,E,τMOpðdÞ  : ð5:117Þ
NDPEðStÞ

F P,Ei ¼ J P,Ei  SE : ð5:117aÞ

The method for determining the membrane element’s standard parameters needed
for this calculation, their extrapolation to the corresponding parameters under design
conditions, and then calculation of the elements’ performance data with regard to
product flow, specific product flux, and product composition follows the sequence as
shown in Table 5.10 for calculating a single RO membrane element.
5.2 RO Membrane Process Design 449

5.2.2.3.2 Composition of Products of Series Element, Module, Stage,


and Array
The composition of product water generated in an array is calculated in steps starting
with determination of the composition of the permeate of each element (Eqs. 5.118–
5.118b) and as also set out in the description of the calculation for a single element in
Sect. 5.1.5.2.5, Table 5.10 for the steps 13 to 16.2.

ciP,Ei ðdÞ ¼ ciFEi ðdÞ  CFEi ðdÞ  βEi ðdÞ  SPiEi ðdÞ : ð5:118Þ

SPiEi ðdÞ ¼ SPiEi ðStÞ  f SPCi : ð5:118aÞ

f TCSðdÞ  f FSi,Ei ,τMOp CFEi ðdÞ  βEi ðdÞ NDPEðStÞ


f SPCi ¼   ð5:118bÞ
f TCWðdÞ  f FP,Ei ,τMOpðdÞ CFiEi ðStÞ  βEðStÞ NDPEi ðdÞ

ciP,Ei ðdÞ ¼ concentration of component i in product of element i at design


conditions [kg/m3; g/l]
SPiEi ðdÞ ¼ salt passage factor of component i at membrane element i under design
conditions []
SPiEi ðStÞ ¼ salt passage factor of component i at membrane element i at standard
test conditions []
f SPCi ¼ salt passage correction factor of component i []
The component concentrations in the products of the membrane modules are then
calculated from the component concentrations in the product water of the membrane
elements (Eq. 5.119). The component concentration ci,P,M i ,Sti ðdÞ in the products of
membrane modules connected in parallel in a single stage could be near or equal to
the concentration ci,P,StiðdÞ in the stage’s product.

PE,M
i¼N PE,M
i¼N
ciP,Ei ðdÞ  F P,EiðdÞ ciP,Ei ðdÞ  F P,EiðdÞ
i¼1 i¼1
ci,P,M i ,Sti ðdÞ ¼ ¼ ci,P,StiðdÞ : ð5:119Þ
PE,M
i¼N F P,M i Sti ðdÞ
F P,Ei ðdÞ
i¼1

ci,P,M i ,Sti ðdÞ ¼ concentration of component i in product of module i in stage i at


design conditions [kg/m3; g/l]
ci,P,StiðdÞ ¼ concentration of component i in product of stage i at design conditions
[kg/m3; g/l]
NE, M ¼ number of elements in module M []
F P,EiðdÞ ¼ product flow of element i in module i at design conditions [m3/h]
F P,Mi Sti ðdÞ ¼ product flow of module i in stage i at design conditions [m3/h]
If the array consists of just one stage as is usually the case for RO systems for
seawater desalination, the composition of the product on module level ci,P,M i ,Sti ðdÞ or
of the stage ci,P,StiðdÞ corresponds to that of the array ci, P, A(d ), i.e. of the RO system.
450 5 Reverse Osmosis Membrane System: Core Process of SWRO

If the array consists of more than one stage, the product composition ci,P,M i ,Sti ðdÞ
of the module Mi,St has to be calculated for each stage Sti, that is its product
concentration ci,P,StiðdÞ corresponding to the flow and concentration conditions
there prevailing (Sect. 5.2.2.3.1). From the component concentrations in the stages
ci,P,Sti ðdÞ , for a multi-stage array the concentrations in the array’s product water ci, P,
A(d ) can then be calculated with Eq. (5.120).

PSt,A
i¼N PSt,A
i¼N
ci,P,Sti ðdÞ  F P,Sti ðdÞ ci,P,Sti ðdÞ  F P,Sti ðdÞ
i¼1 i¼1
ci,P,AðdÞ ¼ ¼ : ð5:120Þ
PSt,A
i¼N F P,AðdÞ
F P,Sti ðdÞ
i¼1

F P,Sti ðdÞ ¼ product flow of stage i at design conditions [m3/h]


ci, P, A(d ) ¼ concentration of component i in product of array at design conditions
[kg/m3; g/l]
FP, A(d ) ¼ product flow of array at design conditions [m3/h]
NSt, A ¼ number of stages in array []

5.2.2.3.3 Method and Sequence of Array Performance Calculation


and Optimization
The feed pressure pF,A to the array that is needed to attain its stipulated product flow
FP,A has to be sufficiently high so that in the final element of the array an adequate
product flow will still be maintained. This is calculated as shown in Eq. (5.121).

J PS,EðN E Þ NDPEðStÞ
pF,A ¼  þ pP,A þ ΔpFC,A þ π F,A
J PS,EðStÞ,V f TCWðdÞ  f FP,E,τMOpðdÞ
 CFF,C,A ðN E Þ  βAðN E Þ  π P,A : ð5:121Þ

pF,A ¼ feed pressure needed for array [bar]


J PS,EðN E Þ ¼ specific product flux of last element of array with NE series elements
[l/m2h]
JPS, E(St), V ¼ specific product flux of last element of array at standard conditions
[l/m2h]
PP,A ¼ product-side pressure of array [bar]
ΔpFC, A ¼ pressure differential of feed to concentrate of array [bar]
CFF,C,S ðN E Þ ¼ concentration factor for feed to brine concentration at last element
NE of array []
βAðN E Þ ¼ concentration polarization factor at last element of array []
π F, A ¼ osmotic pressure in feed to array
π P,A ¼ osmotic pressure in product of array
However, at the start of the calculation process the design parameters of the final
array element needed for calculating the feed pressure, in particular its specific
permeate flux J PS,EðN E Þ and also the concentration polarization coefficient βAðN E Þ ,
5.2 RO Membrane Process Design 451

are not known. Therefore, the calculation is started by taking an approximate value
for the feed pressure as determined in Phase 2 on the basis of the thermodynamic
limit (Eqs. 5.103c, 5.103d and 5.103e). There then follows calculation of each of the
membrane elements in sequence using the set values of Sect. 5.2.2.1, Table 5.11.
After the first calculation run, a check is made of whether the specified product
flow FP,A is attained with the starting value taken for the pressure. If this is not the
case, the pressure must be corrected by changing the input value until the product
flow of the array or of the RO system is attained in accordance with Eq. (5.108).
Other target parameters that have to be complied with or attained during the
calculation run are the guide and limit values stipulated by the membrane manufac-
turer, that are oriented to how the seawater is extracted and treated, so as to enable
dimensioning of each series-connected membrane element and of the array itself (see
Sect. 5.2.2.1 and Table 5.12).
With the selection of the mean product flux JP,S,Ø,A, the numbers of membrane
elements N EA and membrane modules N M A of the array are fixed as shown above in
Phase 2. As the sequential calculation proceeds, however, all other target parameters
given in Table 5.12, like the maximum values of the product flux of the first element
JP,E1, the feed flows FFE to each membrane element as well as the concentrate
discharge FCE and the limiting values of the element recovery rate YE plus the
concentration polarization factor βE of each element, have to be checked and
adjusted.
If after a calculation run these guide values for dimensioning the membrane
elements are not attained, the design parameters of the array, such as number of
membrane elements N EA , number of membrane elements per module NE,M and thus
also the number of membrane elements per array N M A , are adjusted accordingly
following which the calculation is repeated with the changed design parameters. For
a multi-stage array, it may also be necessary to amend the configuration of the stages,
i.e. to increase or reduce the number of membrane modules in the respective stages
N M,Sti . But this adjustment of the structure and configuration of the array will have to
be done such that its prescribed mean product flux JP,S,Ø,A will be retained. To
achieve this it may also be necessary, after adjusting the number of membrane
elements or membrane modules and their allocation to the system’s stages, to correct
the pressure of the feed to the array pF,A. This iterative calculation approach has to
continue until all of the array’s target parameters are attained or they remain within
their prescribed limits.
For these very complex sequential and iterative calculation runs, the membrane
manufacturers provide design tools by means of which, after entering the basic data
listed in Table 5.11 and following selection of a membrane type from one of those
available in the database in the manufacturer’s software together with their
specifications, these calculations can be prepared. The algorithms of these design
programs also are based on calculation equations and procedures comparable to
those set out and described above under Sect. 5.2. However, as a result of the
manufacturers’ specialist know-how and the findings gained by evaluating the
operating outcomes of existing membrane desalination plants, their software
452 5 Reverse Osmosis Membrane System: Core Process of SWRO

includes additional or modified calculation methods that are specific for their
membrane element types, but also incorporate their range of experience in the design
of membrane processes.
The guidelines and limit values as stipulated by the membrane manufacturers for
dimensioning their membrane types are incorporated in all these design tools. If for a
specific design these are violated, the software issues a warning, with the parameters
concerned displayed, stating the element of the array in which violations occur and
their severity. The software user must then manually make the changes to the
configuration of the array as described above, rerun the design calculation, and
keep repeating this procedure until the design values are consistent with the mem-
brane manufacturer’s guidelines. More detailed descriptions of the structure together
with the input and output functions of such design software are provided under
Sect. 5.4.
The results of calculating the performance parameters of a membrane module for
seawater desalination that is equipped with seven membrane elements with a
membrane manufacturer’s13 design program are shown as graphs in Fig. 5.33a–c.
These graphs show how these parameters vary depending on the configuration of the
series-connected elements within the module. The membrane module is supplied
with seawater with a TDS of 35 g/l for a module recovery rate of 45% and at a feed
temperature of 25  C. The membrane elements have a salt rejection under standard
conditions of 99.8% for NaCl and a membrane age of 3 years.
As shown by the graph in Fig. 5.33a, the decrease of the pressure of the feed to the
module’s membrane elements is insignificant in relation to the feed pressure to the
module since the sum of the differential pressures of the unfouled membrane
elements is small in comparison to the pressure applied at the module. As product
recovery increases from element to element with a consequent rise in the concentra-
tion of the feed TDS of the module within the elements, however, there is a
successive increase of the osmotic pressure at their membrane walls, and the net
driving pressure NDPE as the difference of the pressure pFE applied at the respective
element and the osmotic pressure π E,M generated there decrease as the element
number i increases. As a consequence, as shown in the graph of Fig. 5.33b, the
product flow of the series-connected elements deceases from 27% of the module’s
total product recovery at the start to just 5% of the module’s product at the final
element of the series-connected module train.
At the same time, the elements’ salt rejection also drops. For the operating
conditions as stated above and the selected seawater desalination membranes with
a standard salt rejection rate of 99.8%, their effective salt rejection decreases from its
initial value of approximately 99.65% to 98.3% in the module’s final element, i.e. the
TDS of the generated product increases correspondingly from the first to the last
element of the module. The reduction of the product flow of each element is also of
course associated with a reduction of the specific product flux from 24 l/m2.h in the
first element to approximately 4 l/m2.h in the final one (see graph in Fig. 5.33c).

13
Data source: Toray DS2 membrane design program.
5.2 RO Membrane Process Design 453

Pressure [bar]

60

Feed pressure PFE


50
Osmotic pressure πE
at membrane wall of
element Ei
40

Seawater membrane element RE = 99.8 %


Seawater salinity = 35 g/l
30
Recovery = 45%; Temperature = 25°C
Average specific flux = 12 l/m2,h
Membrane age = 3 years

20 Net driving pressure


NDPE
at element Ei

10

0
1 2 3 4 5 6 7
Number of element in membrane module [-]

Rate of recovery RE or
rate of product flow Salt rejection RE of
FPE at element Ei [%) element Ei [%]

50 99.8

45 99.6

40 99.4

35 99.2

30 99.0
Recovery rate Salt rejection RE
YE at element Ei at element Ei
25 98.8

Product flow FPE


20 98.6

15 98.4

10 98.2

5 98.0

0 97.8
1 2 3 4 5 6 7
Number of element in membrane module [-]

Fig. 5.33 (a) RO membrane module—dependence of feed pressure, osmotic pressure, and net
driving pressure on element number i. (b) RO membrane module—dependence of recovery rate,
product flow ratio, and salt rejection on element number i. (c) RO membrane module—dependence
of specific product flux and concentration polarization factor on element number i
454 5 Reverse Osmosis Membrane System: Core Process of SWRO

Specific product flux Concentration


JSPE at element Ei polarisation factor
[l/m2,h] βE at element Ei
25 1.20

1.18

20
1.16

1.14
15
1.12
Concentration polarisation factor
1.10
10
1.08
Specific product flux

1.06
5

1.04

0 1.02
1 2 3 4 5 6 7
Number of element in membrane module [-]

Fig. 5.33 (continued)

The concentration polarization factor βE, that at the start with 1.18 is close to the
maximum permissible value of 1.2, likewise drops to approximately 1.04 at the final
element.
As described above, the performance of a seawater desalination system consisting
of a single-stage array regarding its product flow then results as a corresponding
multiple of the product flow of one module, i.e. the sum of the flows of its elements
(Eq. 5.108), while the product composition of the array is calculated from the
differing product compositions of each of the membrane module’s elements in
accordance with Eq. (5.120).

5.2.2.4 Phase 4: Extension of the Basic Design to Cover the Full Ranges
of TDS, Composition, and Temperature of the Feed
The feed conditions at an RO seawater desalination system may exhibit greater or
lesser fluctuations regarding seawater temperature and TDS depending on local
conditions and seasonal influences. Consequently, during the preparatory
investigations at the early planning stage of a desalination project, profiles of
temperature, TDS, and composition of the seawater are prepared whose variations
provide the basic parameters for designing the membrane separating system (see
Sects. 4.2.1 and 4.2.1.1.1).
This means that it is not sufficient to dimension an RO membrane system for just
one design point, but rather the calculation must model the operating behaviour of
the desalination process under the minimum and the maximum temperature and TDS
conditions that could arise as well as the membrane elements’ change in performance
as they age. Thus, the design parameters for desalination performance, product
5.2 RO Membrane Process Design 455

recovery, and product composition must be held over the entire defined range of
operating conditions and their maximum permissible values may not be exceeded.
Accordingly, membrane calculations for a number of appropriately selected RO
system design cases must be prepared with documentation of the values obtained for
the respective plant operating parameters together with the resulting capacity and
performance data.
If target values for the composition of the product generated in the desalination
system, i.e. TDS, boron, or bromide, are stipulated that cannot be attained by the RO
seawater desalination system on its own, a second pass for post-desalination is
required. For its dimensioning, the bandwidth of the TDS and the concentrations
of limiting components in the product of the upstream first pass have to be consid-
ered (see Sects. 4.2.1.4.1 and 4.2.1.4.2).
With fluctuating feed seawater temperature and TDS, there are two possibilities
for operating an RO system, these being:

• maintaining a constant feed pressure while varying product flow and feed flow
(Operating Option 1), or
• varying the feed pressure while maintaining a constant product flow and feed flow
(Operating Option 2).

For both these options, the product recovery rate, i.e. the ratio of concentrate flow
to feed flow of the RO system, has to be kept more or less constant so that its
membrane elements will not be operated under conditions that lie outside of the
range of the membrane manufacturers’ design guide values. Further, the maximum
product recovery rate selected for system design should not be exceeded, to avoid
scaling and also intensified membrane fouling.
When operating the RO system at a constant feed pressure (Operating Option 1),
the plant’s product flow changes in step with the temperature and TDS fluctuations
in the feed line. The changes in the plant product’s TDS and ionogenic composition
likewise track these feed fluctuations. If for this operating option it is intended that
the product recovery rate is to stay more or less constant, the feed flow to the plant
will then have to be adjusted to match the changing product flow.
With Operating Option 2, by raising or lowering the feed pressure, the
fluctuating water permeability and so product flow of the membrane elements that
arise due to the changing temperature and TDS are compensated. Thus, with this
option the plant can be operated at a uniform product flow and, with a more or less
constant product recovery, also without varying the feed flow.
Normally, commercial, industrial and communal seawater desalination plants are
required to generate a constant product flow unaffected by seasonal temperature
fluctuations and changes in seawater salinity. They are therefore predominantly
operated in accordance with Operating Option 2, which means varying the feed
pressure to maintain a constant product flow and with a product recovery that varies
as little as possible, so factors influencing the product flow are compensated for by
adjusting the feed pressure accordingly. This also applies for the loss of membrane
performance due to their ageing, which is likewise primarily compensated for by
456 5 Reverse Osmosis Membrane System: Core Process of SWRO

raising the feed pressure. Due to varying the feed pressure to a seawater desalination
array of course, also the TDS and ionogenic make-up of its generated desalinated
product are correspondingly influenced.
Therefore, when running the plant in line with Operating Option 2 the pumping
station in the feed line to the RO system must be designed to provide a pressure that
can be varied over a sufficiently wide range.
The following descriptions for expanding the initial design calculations prepared
for a certain design point of a seawater membrane desalination system so that it will
cover the full range of its feed conditions refer to Operating Option 2, that is to
operation of the RO system under varying feed pressure while maintaining a
constant product flow and product recovery.
In order to determine the particular range of values of feed pressure and product
TDS for the framework feed conditions prevailing at the location of a seawater
desalination plant, a series of membrane calculations must be prepared for various
combinations of feed temperature and TDS as well as under consideration of the age
of the selected membrane element. Compiled in Table 5.14 are the corresponding
design cases matched with the required combinations of these design parameters.
These calculations provide an approximate profile of membrane behaviour under
varying feed conditions and with increasing membrane age that can be further
refined by calculating additional operating cases.
The membrane calculations generally commence with the first design case for
which the seawater make-up and feed temperature are determined as the mean of the
minimum and maximum values of these two parameters. The membrane age is the
specified design value. Then with the resulting configuration of the membrane array,
its operating characteristics are modelled under further possible combinations of the
desalination plant’s feed conditions. With the array configuration resulting from the
initial calculation, it should be possible to achieve and also maintain all the perfor-
mance parameters stipulated for the plant within the range of design parameters

Table 5.14 RO membrane design calculation: least required design case calculations
Seawater feed Seawater feed
Design case Membrane age TDS temperature

No. Subject years mg/l C
1 Mean feed pressure Design value Mean of min. Mean of min. and
and product TDS and max. TDS max. temperature
Feed pressure
2 • Maximum Design value Maximum Minimum
3 • Minimum 0 and after Minimum Maximum
membrane
stabilization
Product TDS
4 • Maximum Design value Maximum Maximum
5 • Minimum 0 and after Minimum Minimum
membrane
stabilization
5.2 RO Membrane Process Design 457

stipulated by the membrane manufacturer for the selected membrane element (see
Table 5.12). Should this not be the case, when calculating the array
configuration, then:

• another design point must be taken and the configuration of the membrane array
changed, if necessary with the selection of another membrane type, and if this
modification of the calculation run still does not produce the required result
• the boundary conditions of the desalination feed parameters, i.e. temperature,
TDS and composition of feed, must be reviewed and changed if possible
• a design parameter such as the array’s product flow or its product recovery will
have to be amended.

Particularly for higher product recoveries of between 40% and 50% and a
maximum feed temperature of over 30  C, it may happen that the membrane
manufacturer’s design guide values for the selected membrane element, for instance
the required concentrate-discharge flow, can no longer be maintained and to adjust
the flow through the array accordingly, the target value specified in the design for
product recovery will have to be reduced correspondingly, for example from a YA of
45% to values of 40% to 42%.
The graph in Fig. 5.34 shows for certain bandwidths of temperature and TDS in
the feed line to a seawater desalination array such calculated profiles of the resulting
changes of feed pressure and product TDS.

Feed pressure array pF,A TDS of array


[bar] product [mg/l]

70 900

Seawater membrane element RSt = 99.8 %


Recovery = 45%; 8 elements/module Seawater feed 800
65 Average specific flux = 12 l/m2,h TDS [mg/l]
Membrane age = 3 years
Seawater feed 40,000 700
TDS [mg/l]
60 35,000
40,000 600

55 500
35,000 30,000

400
50

300
30,000
45
200

40 100
10 15 20 25 30 35 40
Sea water feed temperature tF,A [°C]

Fig. 5.34 Seawater RO array design—profiles for feed pressure and product TDS at various values
for feed TDS and temperature
458 5 Reverse Osmosis Membrane System: Core Process of SWRO

These profiles apply for a seawater desalination membrane element with a


standard rejection rate of 99.8% for NaCl that is configured in a module containing
eight elements. The product recovery of the array amounts to 45%, with the average
specific flux stipulated as 12 l/m2.h and the membrane age 3 years. A product
backpressure was not considered for the calculation, i.e. pP is 0. Also, no elevated
differential pressure due to fouling or scaling is considered for the calculated values,
so the figures shown for the feed pressure apply for membranes in clean condition.
For fluctuations in feed TDS, selected for this was a bandwidth of
30,000–40,000 mg/l, and for the temperature 15–35  C. These variations in the
range of feed conditions are greater than would normally be the case for seawater
desalination plants, but these bandwidths are selected as examples to show the order
of magnitude with which changes in the feed conditions at a seawater desalination
plant affect the dimensioning of its components.
From the graph of Fig. 5.34, it can be seen the degree to which the feed pressure
to an array has to be adjusted to maintain the prescribed product flow in order to
accommodate changes to feed TDS and temperature. Thus, to cover the entire
bandwidth of temperature and TDS changes as shown, the feed pressure has to be
adjusted within a range of between around 42 and 59 bar. Thereby the product TDS
is increased from approximately 150 mg/l at the lowest temperature of 15  C and
lowest feed TDS of 30,000 mg/l to over 700 mg/l at the maximum temperature of
35  C and maximum feed TDS of 40,000 mg/l. For cases with lesser changes of the
feed conditions, for example, if the TDS in the feed line varies between 30,000 and
35,000 mg/l for a temperature bandwidth of 20  C to 30  C, the feed pressure must
change between around 42 and 52 bar, while the bandwidth for the change in product
TDS is approximately 220–500 mg/l.
For a design that matches the conditions encountered in practice for RO seawater
desalination, for the chosen membrane elements a certain membrane age is specified
at which a membrane array has to produce the desired product flow. A membrane
element’s permeability decreases with age, although its salt permeability increases.
This means that when the desalination plant commences operation, it can provide a
higher product flow at a lower product TDS than is the case for its design membrane
age (see Sect. 5.1.5.2.3). If the plant is to be operated with the same product flow
from its commencement of operation irrespective of the age of its membrane
elements, its performance must be adjusted accordingly, likewise by changing the
feed pressure, i.e. the net driving pressure.
There thus result corresponding profiles for feed pressure and product TDS for an
RO array, if, for a certain bandwidth, the dependency of these two parameters on the
age of its membrane elements and the temperature of the feed to the array is
calculated as graphed in Fig. 5.35. The dependencies there shown are calculated
for the same membrane configuration and design parameters as for the graph of
Fig. 5.34. The values obtained apply for a TDS in the feed line to the membrane
array of 30,000 mg/l.
The graph in Fig. 5.35 shows, for an unvarying TDS of 30,000 mg/l in the feed
line to the array and with varying temperature, the change of pressure that is needed
5.2 RO Membrane Process Design 459

Feed pressure array pF,A TDS of array product


[bar] [mg/l]

50 600

Seawater membrane element RSt = 99.8 % Average membrane


49 550
Recovery = 45%; 8 elements/module lifetime AMLT
Average specific flux = 12 l/m2,h [year]
48 Average membrane Seawater feed TDS = 30,000 mg/l 500
lifetime AMLT 3
[year] 2
47 450
3 1
46 2 400
0
1
0
45 350

44 300

43 250

42 200

41 150

40 100
10 15 20 25 30 35 40
Sea water feed temperature tF,A [°C]

Fig. 5.35 Seawater RO array design—profiles of feed pressure and product TDS at different RO
membrane AMLT and feed temperature

to attain a constant product flux with increasing membrane age and how the product
TDS of the membrane array then changes.
In order to compensate for the influence of membrane age on product flow, from
the commencement of operation of the membrane array up to when it reaches its
design age of 3 years, for a temperature bandwidth in the feed line of 20  C to 30  C,
feed pressure to the seawater RO has to be adjusted within a range of between around
42 and 45 bar. Under these conditions, the product TDS changes from around
170 mg/l at the start to 380 mg/l.
Combining both profiles as graphed in Figs. 5.34 and 5.35 results in the minimum
and maximum values of feed pressure for which the pumping station has to be
dimensioned as well as the corresponding fluctuation range of product TDS. For the
seawater desalination membrane element selected for calculating the two graphs and
under the defined design criteria, for an array that is supplied with seawater having a
TDS range of 30,000–35,000 mg/l at a temperature fluctuating between 20 and
30  C, a feed pressure bandwidth of approximately 42–52 bar is needed. This feed
pressure bandwidth is made up of the share with which the influence of the varying
temperature and TDS in the feed line to the array is compensated and the signifi-
cantly lower share required to compensate for membrane age.
For these feed conditions, the product TDS lies between 170 and 500 mg/l, from
commencement of operation of the array up to attaining its design membrane age of
3 years. If a second, post-desalination pass is required to maintain a defined product
TDS of the SWRO plant, the capacity of this second RO system has to be dimen-
sioned to provide minimum and maximum product flows as needed to accommodate
460 5 Reverse Osmosis Membrane System: Core Process of SWRO

the fluctuation ranges of the product TDS or other limiting salt constituents emerging
from the first pass.

5.2.3 Options for More Extensive Module, Array, and RO System


Optimization

The basic design for RO membrane systems with spiral-wound element modules as
prepared during the four design phases described above may be optimized, modified,
and adapted through additional technical options to cope with circumstances specific
to the plant. This can be done on the level of configuration of the membrane modules
and of the array as well as for the entire RO system, which in this case would be
based on measures taken for membrane management.

5.2.3.1 Membrane Module Options

5.2.3.1.1 Optimization of Number of Elements per Module


For the design conditions shown in Figs. 5.33a–c, 5.34, and 5.35, it would be
possible to dimension a single-stage array in a module configuration with seven as
well as with eight elements. The advantage of equipping the membrane module with
eight elements is that fewer membrane pressure vessels would then be required to
accommodate the number of membrane elements needed for the array than for the
seven-element configuration (Eq. 5.105). For the selected design case, for the eight-
element module, this would mean around 13% fewer pressure vessels than for a
membrane module with seven elements. Further, the equipment outlay for mem-
brane module pipework and valves within the array is correspondingly reduced.
This is offset by the higher costs for the membrane pressure vessels that are longer
and have a higher pressure loss and thus require a higher feed pressure for the eight-
element module. The reason for the increase in pressure loss is the additional
membrane element, but also the higher feed flow and thus the increased flow through
the eight-element module. The latter is because the total feed flow to the RO array is
distributed over fewer membrane modules (see Eq. 5.111).
Figure 5.36a shows for a typical module equipped with seawater desalination
elements how its total differential pressure varies with the number of membrane
elements with which it is equipped and for increasing product recovery.
However, the rise in differential pressure for an eight-element module compared
with one with fewer elements must be seen in relation to the modules’ feed pressure.
In this case, the increase in the pressure loss in a module with eight elements is of the
order of less than 1% of the feed pressure of a seven-element module.
A comparison of the hydraulic profiles and the performance data between the two
module configurations is shown in the graphs of Fig. 5.36b, c. In these, the curves of
the principal operating parameters of recovery rate YEi, concentration polarization
factor βEi, specific product flux JSPEi, and salt rejection REi are compared for an
individual element i of the membrane module.
5.2 RO Membrane Process Design 461

Differential pressure
of "clean" module
[bar]

1.7
1.6
1.5 Temperature = 20 °C
1.4
1.3
1.2 Membrane elements
per module
1.1
1.0 8
0.9
0.8
0.7 7

0.6 6
0.5
30 32 34 36 38 40 42 44 46 48
Recovery rate of module [%]

Concentration
Recovery rate YE at polarisation factor
element Ei [%] βE at element Ei [-]
20 1.20
19
18 1.18
17
16 1.16
15
1.14
14
13 8 element module
1.12
12 7 element module
11
1.10
10
9 βEi 1.08
8 YEi [%]
7 1.06
6
5 1.04
4
3 1.02
1 2 3 4 5 6 7 8
Number of element in membrane module [-]

Fig. 5.36 (a) RO membrane module—differential pressure of module with its dependence on
number of elements per module and recovery rate. (b) RO membrane module—Comparison of
element recovery rate and concentration polarization factor between a 7-element and an 8-element
module. (c) RO membrane module—Comparison of specific flux and salt rejection between a 7-
element and an 8-element module
462 5 Reverse Osmosis Membrane System: Core Process of SWRO

Specific product flux Salt rejection RE of


JSPE at element Ei element Ei
[l/m2,h] [%]
30 99.8

99.6
25
99.4
RE [ % ]
99.2
20
99.0

15 98.8

98.6
10
98.4
8 element module

7 element module 98.2


5
98.0

0 97.8
1 2 3 4 5 6 7 8
Number of element in membrane module [-]

Fig. 5.36 (continued)

Regarding the recovery rates YEi of the individual elements and their respective
concentration polarizations βEi, it is apparent (Fig. 5.36b) that with the eight-element
configuration, the variation of these parameters is more uniform over the length of
the membrane module than is the case for the seven-element module. With the eight-
element module, the recovery rate and concentration polarization of the frontmost
element are lower than for the seven-element alternative.
The comparison of specific product flux YSPEi and salt rejection REi for the two
module configurations (see Fig. 5.36c) shows that the values for specific product flux
of the eight-element module are somewhat higher, starting with the first membrane
element and continuing over the entire module length for each element in the seven-
element module. The mean of the specific product flux across all elements in the
respective module is, though, the stipulated design value of 12 l/m2h.
Also, for the salt rejection of each element, the eight-element module when
compared with the module with seven elements shows a slightly higher value at
the start but this difference becomes greater with each subsequent element. But here,
too, for the selected example the difference in the mean salt rejection over the entire
module at approximately 99.11% with seven elements compared to 99.09% with
eight elements is only small between the two module types.
The main difference between the two module configurations regarding their
performance profiles is apparent in their fouling characteristics and in the perfor-
mance of the frontmost module elements. For the eight-element module, due to the
comparatively higher specific product flux in the frontmost elements, a somewhat
5.2 RO Membrane Process Design 463

higher propensity for fouling is to be expected (see Sects. 5.3 and 5.3.2). Regarding
the salt rejection of the first elements, a somewhat better product quality may be
expected for “clean” elements for a module with eight elements than is the case for
the seven-element module.

5.2.3.1.2 Hybrid Elements: Internal Staging


In an RO membrane module in which similar membrane elements are connected in
series on the concentrate side, the greatest share of desalinated product water is
generated by the elements at the feed end. The reason for this is that, on account of
the still small increase in concentration there prevailing, the osmotic pressure of the
concentrate at the feed elements in comparison is low while the net driving pressure
there is highest (see the graphs in Fig. 5.33a, b). For the feed conditions, recovery
rate and type of seawater desalination membrane for which the two graphs in these
figures have been calculated; in the first element of this example of a membrane
module equipped with seven membrane elements, approximately 28% and 22% of
the product flow are generated in the first and second elements, respectively, which
account for a total of 50% of the module’s product flow. This means that at the
module’s head end, also the product flux and concentration polarization factor are at
their maximum and these two parameters decrease with element number
(Fig. 5.33c). This unequal distribution in the generation of the desalinated product
has two major disadvantages, these being:

• Particulate and colloidal solids that are still present in the feed line to the module
are deposited more in its first element, which means that fouling is greatest there.
The reason for this is that the accretion of solids at a membrane element obeys the
same physical principles as for the concentration polarization of ionogenic
substances, i.e. the higher the product flux through the membranes is in propor-
tion to the liquid swept along their surfaces, the greater is the rate of deposition on
these surfaces (for more details, see Sects. 5.3 and 5.3.2).
• With a high permeate yield YS of 45% or more and at higher operating
temperatures of greater than 30  C, this uneven distribution of the product flow
is even more pronounced. This means that at elevated temperatures frequently,
the product recovery of an array configuration dimensioned for a lower tempera-
ture has to be reduced to comply with the guide values for the rate of feed to the
elements stipulated by the membrane manufacturer.

A membrane module’s performance with regard to both product flow and product
quality can be appreciably improved if it is equipped with elements with differing
membrane permeabilities and salt rejection rates [66]. With this so-called hybrid or
also “internal staged design” (ISD) configuration of a membrane module, so as to
shift the product flow from the module’s feed end to the downstream elements
towards its discharge end, selected for the first upstream elements are types with a
higher salt rejection and lower product flux (high rejection category) that are then
followed by membrane elements belonging to the category of high product flow with
lower salt rejection rate (high flux/low energy category).
464 5 Reverse Osmosis Membrane System: Core Process of SWRO

Depending on the RO system’s operating conditions, two or more membrane


element categories may be installed in the module to even out performance along the
module. The extent to which the performance at the product end can be improved
over the conventional design depends on how the membrane elements differ with
regard to their membrane permeability and salt rejection characteristics, how many
there are of each type, and how they are distributed in the module. This also applies
to what influence the hybrid, or ISD, design has on the feed pressure to the
membrane module, i.e. whether it remains the same, is reduced, or is raised [67].
The graph of Fig. 5.37 shows a comparison of how the specific product flux and
the salt rejection vary for each element for two eight-element module configurations:
that of a conventional design equipped throughout with the same type of membrane
element and that of a hybrid design with differing membrane types. The module’s
operating conditions with regard to TDS, recovery rate, temperature, and average
specific flux are the same as those shown in Fig. 5.33a. For the two designs in the
ISD configuration, the front elements each have a salt rejection as NaCl of 99.85%
but the downstream elements differ in that one has a rejection rate of 99.80% and the
other 99.75%. This illustrates that the configuration for which the feed end elements
have a high salt rejection, whereas the discharge end elements have a lower salt
rejection but higher product flow show overall the best results regarding total flow
and product quality.

Specific product flux Salt rejection RE of


JSPE at element Ei element Ei
[l/m2,h] [%]

30 100.0

RE [ % ]

25 99.5

20 JSPE [ l/m2,h ] 99.0

15 98.5

10 Similar elements RSt = 99.8 % 98.0

3 front elements RSt = 99.85 %


other elements RSt = 99.80 %
5 97.5
3 front elements RSt = 99.85 %
other elements RSt = 99.75 %

0 97.0
1 2 3 4 5 6 7 8
Number of element in membrane module [-]

Fig. 5.37 RO membrane module—comparison of specific flux and salt rejection between a hybrid
(ISD) design and a conventional element configuration
5.2 RO Membrane Process Design 465

The specific flux of the first element of the module may be reduced from 24 l/m2h
to around 18–19 l/m2h. Thanks to the hybrid configuration, a noticeable improve-
ment is attained in the module’s salt rejection, from a mean of 99.1% to 99.4%.
With this optimization of a membrane module’s element configuration and when
treating problematical raw water with a high fouling potential in particular and also
for a high recovery rate at elevated temperature, an appreciable improvement of the
RO system’s operating conditions is possible. Thus, a single-stage array whose
membrane arrangement with a design temperature of 25  C and other design
conditions as specified in Fig. 5.33a can be operated as a hybrid configuration
over a temperature range of 15  C to 35  C without having to change the recovery
rate. For a conventional configuration with the module equipped with seven or eight
identical elements, as from a temperature of 30  C to 35  C, it may be necessary to
reduce the recovery rate so as not to exceed the elements’ design guide values.
As set out above, a further advantage is the improvement in membrane perfor-
mance regarding product flow and salt rejection.
For raw water with a greater fouling potential in particular, a module configura-
tion with eight elements (see Sect. 5.2.3.1.1) may be advantageously combined with
the hybrid design. As shown by the graph in Fig. 5.37, compared to the conventional
membrane configuration with an eight element module configuration and identical
elements, with a hybrid design a performance profile results for which the specific
product flux of the first element is appreciably reduced and thus its susceptibility to
fouling is lessened. The salt rejection of the front elements, though, is significantly
increased in comparison with the eight-element module with identical elements so
the TDS in the product water generated by these elements shows a further improve-
ment. Another plus point of the eight-element module design is the reduced CAPEX
for the plant.
A drawback of the hybrid configuration is that management of membrane
replacement for attaining a specified membrane age for the entire RO system is
considerably more complex and also there are certain restrictions in the mode of
membrane replacement (see Sect. 5.2.3.3.1)

5.2.3.2 Array Options

5.2.3.2.1 Split Partial Mode


For seawater desalination plants in which potable water is generated, RO systems are
mostly designed as a two-pass configuration in which the product water from the first
pass is fed to the second pass for post-desalination. The product flow from the first
pass is then treated either wholly or in part in the post-desalination stage. How great
the capacity of the second pass is dimensioned and what proportion of the product
flow of the first pass is bypassed around the second pass depends above all on the
composition of the product water after the second pass and that of the permeate from
the first pass (see Sects. 4.2.1.4 and 4.2.1.4.1).
With the so-called split-partial design, the flow for which the second pass RO
stage has to be dimensioned and operated can be optimized. For this, unlike for the
normal second pass partial design, the entire stream of product water generated in the
466 5 Reverse Osmosis Membrane System: Core Process of SWRO

FF,M
cF,M
E1,spl E2,spl E3,spl E4 E5 E6 E7
Feed F Concentrate C

FPE1,spl ciPE1,spl FPE2,spl ciPE2,spl FPE3,spl ciPE3,spl FPE4 ciPE4 FPE5 ciPE5 FPE6 ciPE6 FPE7 cPE7

FP,fr ont ciP,fr ont FP,rear ciP,rear


Product Pfr ont to bypass Product Prear to 2.pass

Element with higher salt rejection


in case of ISD configuration

Fig. 5.38 RO Membrane module with hybrid and split partial configuration

first pass is not piped onto the post-desalination stage, but instead is split into two
parts, one being the front part stream with a lower TDS, i.e. component concentra-
tion, that is extracted from the feed end modules. The rear part stream at a corre-
spondingly higher concentration is taken from the end elements of the moduls. The
front part stream bypasses the post-desalination membrane stage of the second pass,
while the rear part stream is fed to it. Subsequently, the two part streams are
remerged and the blended product should then show the specified composition.
In order to determine the compositions of the two part streams that are already to
be split on membrane module level, how the product flux and concentration of each
component change throughout the module itself have to be known. The flow volume
of the part stream to be branched off, FP,front or FP,rear, and its component concen-
tration ci,P,front or ci,P,rear are calculated from the product flow FPEi and the product
composition ciP,Ei at each membrane element Ei that participate in the generation of
the part stream concerned (Fig. 5.38).
The component concentrations in the front and the rear part streams of the module
result from the product flow and the product composition of each element involved
in generating the part streams, as shown by Eqs. (5.122) and (5.122a).

P
i¼N E,M,F P
i¼N E,M,F
ciP,Ei ,F  F P,Ei,F ciP,Ei ,F  F P,Ei,F
i¼1 i¼1
ci,P,M,F ¼ ¼ : ð5:122Þ
P
i¼N E,M,F F P,M,F
F P,Ei ,F
i¼1

PE,M
i¼N P
i¼N E,M
ciP,Ei ,R  F P,Ei,R ciP,Ei ,R  F P,Ei,R
i¼N E,M,F þ1 i¼N E,M,F þ1
ci,P,M,R ¼ ¼ : ð5:122aÞ
P
i¼N E,M F P,M,R
F P,Ei ,R
i¼N E,M,F þ1

ci, P, M, F ¼ concentration of product of front split flow [kg/m3; g/l]


NE,M,F ¼ number of front elements for front split flow []
ciP,Ei ,F ¼ concentration of product of front element Ei,F [kg/m3; g/l]
F P,Ei ,F ¼ product flow of front element Ei,F [m3/h]
FP,MF ¼ product flow of front split [m3/h]
5.2 RO Membrane Process Design 467

ci, P, M, R ¼ concentration of product of rear split flow [kg/m3; g/l]


ciP,Ei ,R ¼ concentration of product of rear element Ei,R [kg/m3; g/l]
F P,Ei ,R ¼ product flow of rear element Ei,R [m3/h]
FP,MR ¼ product flow of rear split [m3/h]
A graph of the product flux and how the TDS changes for a standard seawater
element with an NaCl rejection of 99.8% under the design conditions of Fig. 5.33a
for a membrane module with eight elements is depicted in Fig. 5.39.
Derived from this graph is the TDS content in the split part streams depending on
how many elements participate in generating the low-salinity front part stream and
then in the rear part stream resulting from this. If the front part stream is generated
with four elements and around 75% of the total product flux of the module, its total
dissolved solids content can be calculated as approximately 180 mg/l. The associated
rear part stream then has a TDS concentration of about 600 mg/l. If the front part
stream is generated from the product of five elements, that is from some 85% of the
module’s product flow, it has a TDS content of approximately 220 mg/l while the
rear part stream one of about 750 mg/l. Without the split partial mode, for the product
of the membrane module results a TDS of approximately 270 mg/l.
The two part streams may be generated in the modules by installing a mechanical
barrier in the product header tube where the front part stream is to be split off from
the rear part stream. The ratio between these two part streams is then fixed and can no
longer be changed [68]. Another possibility is to measure the conductivity in each of
the part stream lines and adjust their flows with control valves. The ratio of the two

Product flow rate at


Product - TDS [mg/l]
element Ei [%]
1,400 100

1,300 SW Standard element RNaCl = 99.8


90
1,200

1,100
80
1,000
Product flow rate
900 70
Rear product TDS with
800
5 elements 60
700

600 4 elements
50
TDS
500
Product TDS without 40
400 split partial

300
30
200 Front product TDS
with 4 or 5 elements

100 20
1 2 3 4 5 6 7 8 9
Number of element in membrane module [-]

Fig. 5.39 Product TDS and flow profile of RO membrane module in split partial configuration
468 5 Reverse Osmosis Membrane System: Core Process of SWRO

1st Pass Seawater RO

E1 E2 E3 E4 E5 E6 E7 E8
Feed Concentrate
E1 E2 E3 E4 E5 E6 E7 E8

E1 E2 E3 E4 E5 E6 E7 E8
2nd Pass Concentrate
Cond Cond
Front Product (low TDS) Rear Product (high
to By-Pass AI TDS) to RO2 AI

CP,E,F ci,P,E,F CP,E,R ci,P,E,R

FI FI

Cond Cond
AI AI

CP,RO2 CP,RM
2nd Pass RO
cP,RO2 cP,RM
Blended Product
VSD
CB,RO2
cB,RO2 By-Pass

Fig. 5.40 Split partial flow scheme

flows is controlled depending on the operating conditions in the second pass and the
composition of the blended product resulting from mixing the bypass flow and the
desalinated product from the second pass, as shown in Fig. 5.40 [69]. As will be
shown below, although this option is more complex, this solution is best suited to the
operating conditions at an RO system in split-partial flow mode.
The flow rate at which the second pass of the RO system has to be operated for
post-desalination so as to attain a defined component concentration or a stipulated
value for the TDS of the generated potable water is characterized by the capacity
factor fC,RO2 (Eq. 5.123). This is calculated from the ratio of the capacity of the RO2
stage CP,RO2 to the production capacity to be generated CP,M, this being equivalent to
the component concentrations in the product in the bypass flow cP,RO2 passing the
RO2 stage to the stipulated component concentrations in the finished product water
cP,M as well as the component concentrations cP,RO2 in the product from the RO2
desalination stage (Eq. 5.123a).

C P,RO2 ¼ f C,RO2  CP,M ð5:123Þ

CP,RO2 c  cP,M
f C,RO2 ¼ ¼ B,RO2 ð5:123aÞ
C P,M cB,RO2  cP,RO2

fC,RO2 ¼ capacity factor RO2 (second RO pass) []


CP,RO2 ¼ capacity of RO2 unit [m3/day]
CP,M ¼ capacity of merged RO product [m3/day]
CB,RO2 ¼ capacity of bypass past the second pass [m3/day]
cB,RO2 ¼ concentration of component of CB,RO2 [mg/l]
cP,M ¼ concentration of component of CP,M [mg/l]
cP,RO2 ¼ concentration of component of CP,RO2 [mg/l]
5.2 RO Membrane Process Design 469

From Eq. (5.123a), it can be seen that fC,RO2 depends significantly on the
concentration of the product cB,RO2 flowing through the bypass line. This explains
the advantage of the split-partial mode for which the concentration of the
components in the bypass line is markedly lower than for the normal partial design.
Likewise from Eq. (5.123a), it can be deduced that fC,RO2 depends both on the salt
rejection of the membrane elements for the components to be separated that in the
seawater desalination system generate the front part stream (RRO1) and on the
rejection (RRO2) of the membrane elements in the post-desalination system RO2
for these components (Eq. 5.123b).

ð1  RRO1 Þ  ccF,RO1
P,M

f C,RO2 ¼ : ð5:123bÞ
ð1  RRO1 Þ  ð1  RRO1 Þ  ð1  RRO2 Þ

cF,RO1 ¼ concentration of component at feed to first pass [mg/l]


RRO1 ¼ salt rejection factor of component of membranes in RO1 []
RRO2 ¼ salt rejection factor of component of membranes in RO2 []
When operating in split-flow mode, the capacity CPEF that has to be generated in
the modules of the first pass to attain a defined capacity CP,RO2 of the post-
desalination pass RO2 is characterized by the split-partial ratio RSP. This is calcu-
lated with Eqs. (5.124) and (5.124a). The lower the capacity CP,RO2 of the post-
desalination stage, the higher is the split-partial ratio RSP, which means the more
front product has to be drawn from the modules of the seawater desalination system
and added to the bypass line of the post-desalination stage.

RSP ¼ 1  f C,RO2 : ð5:124Þ

SPRO1  ccF,RO1
P,M

RSP ¼ 1 
SPRO1  SPRO2  SPRO1
ð1  RRO1 Þ  ccF,RO1
P,M

¼1 : ð5:124aÞ
ð1  RRO1 Þ  ð1  RRO1 Þ  ð1  RRO2 Þ

SPRO1 ¼ salt passage factor of component of membranes in RO1 []


SPRO2 ¼ salt passage factor of component of membranes in RO2 []
RSP ¼ split partial ratio
To obtain a TDS of 100 mg/l in the final product, under the design conditions of
Fig. 5.34, at a feed temperature of 25  C and with a brackish water membrane with
an RNaCl of 95% in the second pass, with a standard partial design a post-desalination
capacity calculates some 74% of the product water capacity factor of the RO system,
whereas with a split-partial design this is only around 58%.
But the temperature of the feed flow to the RO system, too, affects the concentra-
tion conditions in the post-desalination stage and greatly influences the capacity
factor of the RO process in the second pass.
470 5 Reverse Osmosis Membrane System: Core Process of SWRO

Capacity factor RO2


Split partial rate RSP [%] fC,RO2 [%]
75 100
70 RO feed TDS = 35,000 mg/l
Final product TDS = 100 mg/l 90
65
Membrane 2nd Pass RNaCl = 99.5 %
60
55 80

50 RSP fC,RO2
70
45
40
60
35
30
50
25
20 40
15
Membranes 1st Pass
10 High rejection RNaCl = 99.85 % 30
5 Standard RNaCl = 99.80 %
Low energy RNaCl = 99.75 %
0 20
15 20 25 30 35
RO feed temperature [°C]

Capacity factor RO2


fC,RO2 [%]
100

Membrane 1st Pass : Standard SW RNaCl = 99.80 %


90 Membrane 2nd Pass : Standard BW RNaCl = 99.50 %

80

70

60
Feed TDS [mg/l]

50 45,000

40,000
40
35,000
30
30,000

20
10.0 15.0 20.0 25.0 30.0 35.0 40.0
RO feed temperature [°C]

Fig. 5.41 (a) Two-pass split-partial design—dependence of split-partial rate and second-pass
capacity on type of membrane and RO feed temperature. (b) Two-pass split-partial design—
dependence of second pass capacity on first pass feed TDS and feed temperature
5.2 RO Membrane Process Design 471

For a specified target value of the salinity in the product water with a TDS of
100 mg/l, Fig. 5.41a shows how the capacity factor fC,RO2 and the associated split-
partial ratio RSP depend on the type of separation membranes used in the seawater
desalination stage as well as on the temperature of the feed flow to the RO system.
For seawater membranes in the first pass with high salt rejection when compared
with membranes with a lower rejection, it is found that the required capacity of the
RO post-desalination stage can be much reduced. For the design conditions as shown
in Fig. 5.41a, this corresponds to a difference in the RO2 capacity from 10% up to
25%.
Temperature variations in the feed line to the RO system in turn result in
corresponding increases or reductions of the component concentrations in the first
stage product and also a changed salt rejection in the second stage, whose opera-
tional capacity factor has to be adjusted so as to maintain the specified target value in
the finished product.
For a standard seawater RO membrane in the first stage, for temperature
fluctuations in a range of between 20  C and 30  C in the feed line to the RO
system, it must be possible to vary the capacity factor of the second pass between
around 45% and 70% of the desalination plant’s product flow. For very high
seasonal temperature differences at the location of a seawater desalination plant,
this could mean that the capacity of its second pass will even have to be capable of
being varied by more than 50% of the capacity of the SWRO plant.
Shown in Fig. 5.41b is how changes in the TDS and the temperature in the feed
line to a desalination plant impact the capacity needed in its second pass. Tempera-
ture and TDS have their greatest influence in the low temperature range of between
15  C and 20  C. Here, a change of the feed TDS by 5000 mg/l can result in a change
in the capacity factor in the second pass of between 10% and 15%. Regarding its
capacity range, the post-desalination stage, however, must also exhibit sufficient
flexibility to compensate not only for variations of TDS and temperature in the feed
line to the seawater RO stage, but also for deterioration in its salt rejection
characteristics due to fouling.
The operating conditions described above show that the second pass of a seawater
desalination plant for generating potable water must display a high degree of
flexibility in its capacity range so as to compensate for variations in the feed
conditions and also to cope with operational changes in the performance of the
seawater desalination section. This means that, with a split-partial design of the post-
desalination stage, a fixed, defined setting of the ratio of the front part stream to the
rear part stream at the seawater RO membrane modules would not be capable of
fulfilling these requirements for variability of the RO systems capacity. When
designing the second pass, the flexibility requirements of the post-desalination
stage’s capacity have to be taken into account, like providing variable speed control
of the RO2 feed pumps and through a modular configuration of the RO2 system, for
example in the number and capacities of its trains.
The capacity of the second pass must be designed to cater for the operating
conditions for which its factor fC,RO2 is a maximum. Depending on the plant’s
location-specific feed conditions, this could be the combination of “lowest tempera-
ture—highest TDS” or “highest temperature—highest TDS”.
472 5 Reverse Osmosis Membrane System: Core Process of SWRO

When designing the second pass in the split-partial configuration, it must be


borne in mind that the results of calculating a membrane system with the aid of a
membrane manufacturer’s design software refer to “pristine” membranes, that is to
ideal operating conditions with the influence of fouling and scaling on the membrane
elements’ capacity and salt rejection not fully taken into account.
Experience shows, though, that the front elements of a membrane module are the
most prone to fouling. Their performance can accordingly deteriorate more rapidly
than is the case for the other membrane elements. Under this aspect, too, the
combination of split-partial design with a hybrid configuration of the membrane
modules is advantageous: thanks to the membrane elements with higher salt rejec-
tion being installed at the front, not only is it primarily the quality of the module’s
front part stream that is improved, but, as described above, also the specific flux is
reduced here. As experience at existing plants has shown, this has a favourable effect
on the front modules’ fouling characteristics and on membrane lifetime. But in any
case, when dimensioning the capacity of the second pass RO stage, an additional
safety margin should be allowed to ensure that, even over extended operating
periods under possibly arduous operating conditions, the post-desalination stage
will be able to reliably fulfil its task.

5.2.3.2.2 Product Throttling


The capacity of an RO membrane system is matched to variable operating
conditions, like fluctuations of temperature or TDS in the feed line as well as
reductions of the membranes’ water permeability due to ageing or fouling, by raising
or lowering the net driving pressure (NDP) at the membrane elements. The NDP can
be varied by changing the feed pressure to the membrane system and also by
increasing the pressure on the membranes’ product side (see Eq. 5.113). This
product-side pressure increase is done by installing throttling valves in the product
line of the RO stage or in the permeate lines of the array and throttling the respective
product flow and is therefore termed product or permeate throttling. However, this
measure for NDP adjustment is only possible for changes in operating conditions
that lead to an increase in membrane permeability, such as an increase in temperature
or a decrease in salinity in the RO feed.
During RO operation, i.e. under dynamic operating conditions, it is possible for a
backpressure to be generated on the membranes’ product side that exceeds the
product pressure that is permissible under static conditions (see Sect. 5.1.3.1).
When the plant is not operating, however, the technical design must ensure that
the maximum permissible permeate backpressure will not be exceeded, so appropri-
ate pressure relief has to be provided for this case.
For multi-stage arrays, product throttling enables the product flow to be balanced
for the individual stages, i.e. their respective net driving pressure and thus also their
product flow is adjusted by throttling so that the product output of the array is
distributed more evenly over all stages. This may be done alternatively by means of
an interstage booster pump. Product throttling can be done with less operative outlay
and lower capital costs compared with such a booster pump, but at the cost of higher
energy consumption.
5.2 RO Membrane Process Design 473

The aim when operating RO systems is to attain a mostly constant product flow,
which is done by changing the feed pressure as little as possible irrespective of the
varying operating conditions in the plant’s feed as well as of how membrane ageing
degrades the permeability of the installed membrane elements. This results in a
further application for product throttling; this being to compensate for the influence
of membrane ageing as well as of temperature and TDS fluctuations on the mem-
brane elements’ water permeability and thus on the array’s product flow. This
application is particularly of interest for RO plants whose feed conditions show
broad bandwidths of temperature and TDS. In this case, through additional product
throttling, the outlay for pressure control in the desalination plant’s feed can be
reduced. But then, due to the throttling on the product side, there is an increase in
energy consumption.
At the commencement of operation of an RO plant, its new membrane elements
exhibit higher water permeability than is the case when the membranes approach the
end of their design life. So as to be able to operate the plant at the same capacity right
up to the end of its design membrane age, either:

• it can be started up at a reduced feed pressure that can then be progressively


raised, or
• it can be operated from the start at the feed pressure that will be required when it
attains the design membrane age, but the net driving pressure will then be raised
by product throttling, with a progressive reduction of the degree of this throttling
during operation up to the end of the specified membrane service life.

With product throttling, the backpressure pP,A needed on the product side
depends on how the membrane ages, i.e. on how the permeability decline
factor f FP,E,τMOp changes over time (see Sect. 5.1.5.2.3). The design membrane age
τOp and the resulting permeability decline factor f FP,E,τMOp determine the product
backpressure that will have to be applied at the commencement of plant operation
and the degree to which it will have to be reduced up to the end of the specified
membrane lifetime so as to maintain a constant feed pressure pF,A to the array.
The product backpressure, though, is additionally influenced by temperature, that
is the variation with time of the temperature correction factor fTCW for the
membranes’ water permeability, as well as the TDS, which effectively determines
the osmotic pressure π F, A in the feed line. If there is a temperature increase in the
feed line and/or a reduction in TDS, the resulting increase in membrane permeabil-
ity, and thus an elevated product flow, can likewise be compensated by product
throttling. In this case, the product backpressure will have to be increased corre-
spondingly beyond the value needed for compensating for membrane ageing. Per-
meability decline will be compensated by raising the feed pressure.
Equations for calculating the product backpressure pP,A needed for product
throttling may be derived from Eq. (5.121), i.e. the equation for determining the
feed pressure pF,A for an RO array. If the water permeability coefficient of the
membrane elements A(St), V, E under standard conditions is known, the product
474 5 Reverse Osmosis Membrane System: Core Process of SWRO

backpressure pP,A is estimated as shown in Eq. (5.125). Alternatively, Eq. (5.125a) is


applied if the specific product flux JPA, E(St), (NE) and the net driving pressure NDP(St),
E under standard conditions of the elements are known as they result from the
manufacturers’ membrane specifications.

f FP,E,τMOp J PðtÞA,EðN E Þ
pP,A ¼ pF,A  Δπ C,m,P,AðN E Þ   : ð5:125Þ
f TCW AðStÞ,V,E

Δπ C,m,P,AðN E Þ ¼ π FðtÞ,A  CFF,C,A ðN E Þ  βAðN E Þ  π P,A :

J PðtÞA,EðN E Þ f FP,E,τMOp
pP,A ¼ pF,A  Δπ C,m,P,AðN E Þ   NDPðStÞ,E  : ð5:125aÞ
J PA,EðStÞ,ðN E Þ f TCW

pF,A ¼ feed pressure needed for array [bar]


pP,A ¼ product pressure for stage or array [bar]
Δπ C,m,P,AðN E Þ ¼ osmotic pressure difference between concentrate at membrane
wall of final element of array with NE series-connected elements and osmotic
pressure of product [bar]
J PðtÞA,EðN E Þ ¼ specific product flux of final element of stage or array with NE
series-connected elements [l/m2h]
J PA,EðStÞ,ðN E Þ ¼ specific product flux of final element of stage or array with NE
serial elements at standard conditions [l/m2h]
ΔpFC, A ¼ pressure differential of feed to concentrate of stage or array [bar]
CFF,C,AðN E Þ ¼ concentration factor for feed to brine concentration at final element
NE of stage or array []
βAðN E Þ ¼ concentration polarization factor at final element of stage or array []
The type of NDP control is already to be fixed during the design of a seawater
desalination system, with a decision to be taken on basis of a cost comparison either
in favour of just feed-side pressure control or additionally throttling on the
product side.

5.2.3.2.3 High Rejection SWRO and Brine Recovery


The recovery rate that can be attained with an RO membrane desalination process for
a certain seawater TDS is determined and influenced by a number of parameters, for
example:

• the thermodynamic limit, i.e. the limiting pressure that at a certain recovery rate
and the then existing TDS of the RO concentrate is needed to compensate the
osmotic pressure which then is prevailing there (see Sect. 5.2.2.2.1, Eq. 5.103c
and Fig. 5.32)
• the membrane elements that are selected for a RO unit, their permeability, and salt
rejection characteristics as well as the maximum operating pressure they can
withstand
5.2 RO Membrane Process Design 475

• the configuration of the membrane elements in the modules and how in turn these
are arranged in the array to minimize concentration polarization and comply with
the design criteria of the membrane manufacturers
• the corresponding values of the NDP and how these vary within the membrane
modules and the array to attain, with the selected membrane configuration, the
specified product flow through the membranes, and the desired salt rejection
• the hydraulic pressure losses arising within the RO system’s plant configuration
under its operating conditions.

Having a substantial influence on the attainable recovery rate and the feed
pressure at an RO membrane system needed for this purpose is the distribution
and variation of the NDP within the membrane array and, particularly in the region
of high TDS and osmotic pressure, the NDP still available there.
With a seawater desalination system of conventional design as a single-stage
array, the NDP progressively drops to reach a minimum in the last element due to the
decreasing feed pressure to the membrane elements across the membrane module
and the simultaneously increasing osmotic pressure with increasing number of
elements. The consequence is that the major share of product flow is generated in
the front elements of the membrane module and of the array (see Sect. 5.2.2.3.2 and
Fig. 5.33a–c).
A marked improvement over this design is the hybrid configuration, i.e. the
internally staged design, ISD (see Sects. 5.2.3.1 and 5.2.3.1.2 and Fig. 5.37),
which more evenly balances the course of the NPD in the membrane module and
reduces the degree of reduction of its values in the elements from the first to the last
element of the module
A further design possibility is to split the RO array into several stages and, by
means of booster pumps between these, raise the feed pressure to each subsequent
stage depending on the degree of concentration increase and the resulting osmotic
pressure there prevailing. In this way, the NDP can be adapted much better to the
increase in the osmotic pressure, the flow through the reverse osmosis array can be
made more balanced, and overall higher concentration increase and thus greater
product recovery can be achieved.
Such a seawater desalination design with a two-stage array in which the concen-
trate of the first stage is fed to a second stage with a booster pump is shown in the
process schematic in Fig. 5.42. The concentrate of the second stage is fed to an
energy recovery system located in this stage which can completely provide the
power demand of the booster pump from the pressure still existing in the concentrate
of the second stage. Any excess power not required for this purpose is supplied to the
HP feed pump of the array.
The configuration shown in Fig. 5.42 was trialled in a pilot plant for the first time
by the membrane manufacturer Toray commencing in 1997. The maximum recovery
rate possible with this configuration under specific seawater feed conditions of TDS
and temperature range depends essentially on the maximum operating pressure that
the membrane elements installed in the array’s second stage can withstand. For this
purpose, Toray developed a special HP membrane with a salt (NaCl) rejection rate of
476 5 Reverse Osmosis Membrane System: Core Process of SWRO

Seawater RO array
Y ≈ 50 - 60%
HP feed
pump
Seawater feed 1st Stage Product water 1st stage
Y ≈ 35% - 45%

Concentrate
1st stage Booster
pump
2nd Stage
Y ≈ 20% - 35%
Product water Product water
2nd stage of array
Concentrate
2nd stage
Energy recovery

Concentrate
of array

Fig. 5.42 High recovery seawater RO array with two stages and booster pump (brine recovery)

99.8% that could be operated at an operating pressure of up to 100 bar. The second
stage of this desalination configuration was equipped with this membrane, while its
first stage contained standard seawater desalination membranes [70].
Compared with the conventional design of an RO seawater desalination plant as a
single-stage array, for a configuration with a two-stage array and a booster pump the
pressure in the feed line to the second stage is adjusted depending on the osmotic
pressure arising there at the specified product recovery rate as well as the permeabil-
ity of the membrane elements installed there (see Fig. 5.43a).
In this way, the NDP which drops to its lowest value in the final element of the
first stage is elevated again in the feed line to the second stage so that it is better
matched to the osmotic pressure that rises throughout this stage. Depending on the
seawater TDS and the temperature bandwidth in the feed to the desalination plant,
the booster pump will have to raise the pressure of the concentrate from the first stage
by 10–20 bar. Because of the variability of the feed to the desalination plant
regarding TDS and temperature, it must be possible to adjust the booster pump’s
discharge pressure accordingly. Thus, to cater for greatly fluctuating seawater TDS
and temperature, both the HP feed pump for the first RO stage and the second-stage
booster pump should be of the variable speed type to control the respective feed
pressures.
From 1997 to 2000, a number of seawater desalination plants with this configu-
ration, designated by Toray as “brine conversion system (BCS)”, came on stream in
the Caribbean and in Spain. Several desalination systems of this type were conceived
right from the outset as BCS plants and then constructed as such. At other plants,
existing conventional SWROs were retrofitted with a second stage for brine conver-
sion to raise their product water capacities without having to expand the pretreatment
capacity accordingly due to the higher product recovery of the BCS system. This
advantage of the BCS configuration has been exploited at a number of Spanish plants
in Maspalomas, Las Palmas, and Ibiza [71–74].
5.2 RO Membrane Process Design 477

Pressure [bar]

90

80
Osmotic pressure
70 Feed πE at membrane
pressure PFE wall of element Ei

60

50 Seawater membrane elements


1st stage RE = 99.75 %
2nd stage RE = 99.85 %
40 Seawater salinity = 35 g/l
Recovery = 60%; Temperature = 25°C
Average specific flux = 12 l/m2,h
30 Membrane age = 3 years
Net driving
pressure NDPE
20
at element Ei

10

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Number of serial element in array [-]

Specific flux [l/m2,h] Salt rejection [%]


24 99.80

22 99.70
Salt rejection RE
at element Ei 99.60
20
99.50
18
99.40
16
99.30
14
99.20
12
99.10
10
99.00
8 Specific flux JE,SP at
element Ei 98.90

6 98.80

4 98.70
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Number of serial element in array [-]

Fig. 5.43 (a) Variation of feed pressure, osmotic pressure, and net driving pressure over the
elements in a two-stage RO seawater system with booster pump. (b) Variation of specific flux and
salt rejection at the elements in a two-stage RO seawater system with booster pump
478 5 Reverse Osmosis Membrane System: Core Process of SWRO

For seawater salinities of 35,000–39,000 mg/l, recovery rates of up to 60% have


been attained with BCS plants. The recovery rate of the first stage was around 40%
and that of the second stage between 25% and 30%. The first stage of the BCS plant
was run at an operating pressure of 57–68 bar and the second stage in a range from
70 bar up to 90 bar with the product TDS of the array quoted as 200–400 mg/l,
depending on membrane age and seawater temperature. The product TDS in the
array’s BCS stage was not all that much higher than that of the first stage. The
explanation for this is that, due to the low recovery rate in the array’s second stage,
also its concentration polarization is correspondingly lower than in the membrane
elements at the end of the first stage. Thanks to the elevation of the NDP in the feed
line to the second stage, the specific flux there also increases accordingly with a
corresponding positive affect on the salt rejection of its membrane elements (see
Fig. 5.43b). Together with the lower salt permeability of the membrane elements of
the second stage, there results a salt rejection of the elements which, despite the
prevailing higher TDS concentration, generates a product quality that is comparable
with the product TDS of the system’s first stage.
Toray has been awarded a number of international patents for its brine conversion
system, among others, in the USA and the EU [75, 76].
With the spiral-wound membrane elements available today and at their maximum
operating pressure of 83 bar, using the membrane manufacturers’ design software,
depending on the seawater’s TDS and temperature, already potential recovery rates
approaching 50% and more can be calculated for a single-stage array in a conven-
tional configuration or with a hybrid design. As compiled in Table 5.15, seawater
TDS of between 35,000 and 45,000 mg/l and a temperature range of between 15  C
and 35  C are the maximum recovery rates obtained for the following RO array
configurations:

• single-stage with identical membrane elements and modules with eight elements
per module
• single-stage in a hybrid configuration with seven elements per module
• two-stage with identical membrane elements, seven elements per module and
inter-stage booster pumps.

These calculations were prepared using membrane design software from various
membrane manufacturers14 with in each case comparable membrane elements from
the various manufacturers for the configuration to be calculated. Degradation of the
recovery rate due to fouling and scaling has not been considered.
The maximum product recovery values in bold in Table 5.15 are determined by
the maximum allowable operating pressure specified by the respective membrane
manufacturers for their seawater reverse osmosis membranes. Taken into consider-
ation and depending on membrane manufacturer was the extent to which they

14
Filmtec ROSA Version 9.1, Hydranautics IMSDesign 2015, LG NanoH20 Q+ Version 2.3, Toray
DS2.
5.2 RO Membrane Process Design 479

Table 5.15 Maximum recovery rates with different array configurations


RO array configuration
Single-
Single-stage— stage— Two-stage with
Seawater Temperature identical elements hybrid booster pump
TDS [mg/l] ( C) Max. attainable recovery (%)
35,000 15 51–59 51–56 59–65
25 49–59 57–61
30 45–59 59–60
35 43–56 50–56 56–58
40,000 15 52–59 51–56 56–61
25 47–59 54–56
30 45–57 52–55–56
35 43–54 49–56 50–52–54
45,000 15 51–53 48–50–56 50–54
25 47–52 51–51–56 49–53
30 45–52 50–50–55 49–51–51
35 43–49 47–48 48–49–49
Design conditions
Type of extraction and Open intake and conventional pretreatment
pretreatment:
Mean product flux: 12 l/m2h
Membrane age: 3 years
Installed membrane elements and their configuration
Conventional design: High rejection elements;
8 elements/module
Hybrid design: First 3 elements: high rejection
Subsequent elements: low energy;
8 elements/module
Brine recovery design: First stage: low energy
Second stage: high rejection
7 elements/module

permitted operation of the membrane elements at their maximum operating pressure


of 83 bar or whether, with increasing temperature, they reduced this pressure
accordingly (see Sects. 5.1.3.1 and 5.1.3.1.1, Fig. 5.13). For the other maximum
recovery rate values quoted in the table, it is design guidelines or limits specified by
the membrane manufacturer, such as the maximum permissible product flux or
maximum permissible recovery rates of the individual elements that limit the
recovery rate before attaining the membranes’ maximum permissible operating
pressure. The explanation for the range of these maximum recovery rates is in part
that, depending on their experience but also on their safety philosophy, the mem-
brane manufacturers set differing limits for their guidelines for designing a mem-
brane system (see Sects. 5.2.2 and 5.2.2.1, Table 5.12). This applies in particular for
the maximum permissible product flux or recovery rate of an element. The higher
this value dependend on type of extraction and pretreatment of the seawater is
480 5 Reverse Osmosis Membrane System: Core Process of SWRO

specified, the nearer the system’s calculated recovery rate can approach the limit
imposed by the membrane element’s maximum operating pressure.This means that,
when designing a seawater RO desalination system, its attainable recovery rate does
not depend solely on the maximum permissible operating pressure of the respective
membrane element, but, in the range of low to moderate seawater TDS, above all on
the manufacturer’s design guidelines, whose limits in turn are stipulated for the
purpose of adherence to specific concentration polarization values as well as for
minimization of fouling.
When comparing the tabulated values for the array configuration in a single-stage
conventional design with that of the hybrid design, one advantage of the hybrid
option becomes apparent in that variations in TDS and temperature have much less
of an impact on the potential maximum recovery rate than is the case for the
conventional configuration.

5.2.3.3 System Options

5.2.3.3.1 Membrane Element Age, Average Membrane Lifetime,


and Replacement Rate of Membranes
The age structure of the membrane elements is a decisive factor for the design and
operation of an RO system and has a major influence on its capacity and salt
rejection as well as its energy consumption (see Sect. 5.1.5.2.3). During operation
of the desalination plant, the membrane elements in an RO array, or rather in its
membrane modules, are susceptible to natural ageing and fouling to varying degrees,
and as a result, their longevities will diverge. Thus, the front elements of a membrane
module have to be replaced more frequently than the downstream elements. Conse-
quently, after a plant has been operating for 1 or 2 years, an array will contain
membrane elements with greatly differing projected operational lifetimes. The
performance and salt rejection of the array as well as the operating pressure needed
to compensate for membrane ageing are derived, for a specified operating time of the
system, from its average membrane lifetime (AMLT) τMOp, ∅ , S. With increasing
operating time, the AMLT value within the system rises although this is mitigated
due to routine replacement of ageing membrane elements by new ones. The ratio of
the initial membrane elements remaining from the original system installation N Eini
to the newly installed elements N Ei drops steadily until a point in time τSOP, equi is
reached when all the original elements have been replaced. The point in time when
this happens depends on the value of the membrane element replacement rate
factor f RRM,i or on the percentage membrane element replacement rate RRM, i,
i.e. the ratio of newly installed elements N Ei to the total number of membrane
elements in the system N ES (Eqs. 5.126, 5.126a, and 5.126b) and the point in time
τSOP,M R,Ei of each membrane replacement operation.
5.2 RO Membrane Process Design 481

N Ei
f RRM,i ¼ : ð5:126Þ
N ES
N Ei  100
RRM,i ¼ : ð5:126aÞ
N ES
N ES  RRM,i
N Ei ¼ N ES  f RRM,i ¼ : ð5:126bÞ
100
f RRM,i ¼ membrane element replacement rate factor at i ¼ 1 year to i ¼ τSOP, AMLT
[]
RRM, i ¼ membrane element replacement rate at i ¼ 1 year to i ¼ τSOP, AMLT [%]
N Ei ¼ number of membrane elements of τMOp,Ei from i ¼ 1 to i ¼ τSOP, AMLT []
N ES ¼ number of membrane elements in the membrane system []
The time at which all of the initially installed membrane elements have been
replaced is termed the steady-state operating point τSOP, equil. At this time, the AMLT
attains its maximum value corresponding to the membrane element replacement rate
RRM, i. This value can then at a constant membrane exchange rate be maintained
over the further operating time of the plant.
In step with the increasing AMLT value, the operating pressure needed to
compensate for membrane ageing and to maintain the RO capacity also rises (see
Sect. 5.2.2.4 and Fig. 5.35). This, too, reaches its maximum value for “clean”
membranes at its steady-state operating point and remains at this level as long as
the membrane exchange rate stays the same throughout the system’s further opera-
tional lifetime.
Upon the RO system reaching a certain operating time τSOP, AMLT, the value of
the AMLT is calculated as the quotient of its entire membrane age τMOp, S at this
point in time and the total number of membrane elements then installed N ES
(Eq. 5.127).

τMOp,S
τMOp,∅,S¼AMLT¼ : ð5:127Þ
N ES

τMOp, ∅ , S¼ AMLT ¼ average membrane lifetime of system at τSOP, AMLT [years]


τMOp, S ¼ membrane age of system at τSOP, AMLT [years]
The system membrane age τMOp, S, in turn, is calculated from the membrane
lifetime of each element τMOp,Ei and the proportion of elements N Ei with the same
lifetime τMOp,Ei in the total number of membranes in the system N ES plus the number
of elements N Eini remaining from the original installation and their membrane
lifetime (Eq. 5.127a).

P 1
i¼τSOP,AMLT
N Ei  τMOp,Ei þ N Eini  τSOP,AMLT
i¼1
τMOp,S ¼ ð5:127aÞ
N ES
482 5 Reverse Osmosis Membrane System: Core Process of SWRO

τMOp,Ei ¼ membrane lifetime of elements for τMOp,Ei ¼ 1 year to τMOp,Ei ¼ τSOP,


AMLT [years]
N Eini ¼ number of initial membranes at τSOP, AMLT []
τSOP, AMLT ¼ RO system operation time at AMLT calculation [years]
The membrane lifetime of a single element is calculated using Eq. (5.127b) from
the operating time τSOP, AMLT of the plant for which the AMLT value is to be
calculated and the time at which the membrane element was replaced.
 
τMOp,Ei = τSOP,AMLT  τSOP,M R,Ei þ 1: ð5:127bÞ

τSOP,M R,Ei ¼ RO system operation time at replacement of membrane element


i [years]
The number of original membranes still remaining after the last membrane
replacement operation N Eini is calculated from the total number of elements in the
system N ES minus the elements that were replaced at the time of the AMLT
calculation N Ei .

X 1
i¼τSOP,AMLT
N Eini ¼ N ES  N Ei : ð5:127cÞ
i¼1

The algorithms of Eqs. (5.127a)–(5.127c) apply if the membrane age of the RO


system or the AMLT is calculated at the end of the respective operating period
following a membrane replacement action.
The design calculation of a membrane system is based either on a value for the
age of the system’s membrane elementsτMOp, ∅ , S or on a membrane replacement
rate RRM. Both these mutually dependent values result from a membrane replace-
ment scenario that has to be prepared as the basis for the design of the system
together with assumptions for the RO plant’s operating period τSOp, i at which time a
start has to be made with membrane replacement and which annual replacement rate
RRM, i is then chosen for this purpose.
Table 5.16 shows a membrane replacement schedule with the help of which the
required membrane age or membrane replacement rate can be determined upon
specifying the value for one of the two parameters which then fixes the other.
The value for the membrane age that is the basis for the RO system’s design
calculation is the AMLT at the steady-state operating point τSOP, equil. For a mem-
brane replacement rate of 13% and commencement of membrane replacement after
an operating period of 2 years, as shown in Table 5.16, the membrane replacement
steady-state operating point will be attained after an 8- to 9-year operating period,
and at this time, the AMLT will have a value of 4.5. Thus, the input data for
membrane design under the conditions of Table 5.16 are a membrane age of
4.5 years and a membrane replacement rate of 13%.
5.2 RO Membrane Process Design 483

Table 5.16 RO system—membrane replacement schedule and resulting lifetime of membrane


elements and average membrane lifetime AMLT of system
RO system years 1 2 3 4 5 6 7 8 9 10
operation time
τSOP,i
Annual %/ 0 0 13 13 13 13 13 13 13 13
membrane year
replacement
rate RRM,i
Membrane years Percentage/number Nei of membranes with τ MOp,E,i [%, No.]
element lifetime 1 100 0 13 13 13 13 13 13 13 13
τMOp,E,i 2 100 0 13 13 13 13 13 13 13
3 87 0 13 13 13 13 13 13
4 74 0 13 13 13 13 13
5 61 0 13 13 13 13
6 48 0 13 13 13
7 35 0 13 13
8 22 0 9
9 9 0
10 0
Average membrane age AMLT
Course of years 1.0 2.0 2.7 3.4 3.8 4.2 4.4 4.5 4.5 4.5
AMLT
AMLT at 4.5
steady-state
Average of 3.4
AMLT till
steady state
Replacement Scheme
Cumulative % 0 0 13 26 39 52 65 78 91 100
replacement
rate initial
membranes
Average 0 0 4.3 6.5 7.8 8.7 9.3 9.8 10.1 10.4
annual
replacement
rate

The dependency of the average membrane lifetime AMLT on differing values of the
membrane replacement rate RRMi in the range from 10% to 20%, as is usual for seawater
desalination systems and after which operating period the plant attains the steady-state
operating point, is shown in Fig. 5.44. The calculations for this graph assume that
membranes will not be replaced until after a system operating period of 2 years.
The AMLT value is stated with reference to the plant operating period prior to the
membrane changeout concerned.
With a membrane element replacement rate of 10%, the steady-state operating
point will be attained after an operating period of around 10 years and an associated
AMLT of 5.6 years, but with a replacement rate of 20% this operating period and
AMLT will be only 5 and 3.2 years, respectively. A membrane element replacement
484 5 Reverse Osmosis Membrane System: Core Process of SWRO

Average membrane
lifetime AMLT [years]
7

Membrane replacement
6 rate RRM [%]
10

13

4 15

20
3

1
1 2 3 4 5 6 7 8 9 10 11 12
Operation time RO system τSOP [years]

Fig. 5.44 RO system—membrane lifetime versus system operation time at various membrane
element replacement rates

rate of 15% corresponds to an AMLT of 4 years and steady-state condition after


7 years.
If the system’s AMLT is to be kept at the same level upon reaching the steady-
state operating point over the RO plant’s remaining operational lifetime, the mem-
brane replacement rate will have to continue at the same level as before the onset of
the steady-state operational phase.
Specifying the membrane age for the design calculation of an RO system fixes at
the same time the membrane element replacement rate RRM. The value of RRM
provides the basis for calculating the annual costs for the membranes and their share
in the desalination plant’s total operating costs. As shown by the graph in Fig. 5.44,
as membrane age increases, the membrane replacement rate drops, so by changing
this parameter the share of the cost of the membranes in an RO plant’s operating
costs can be influenced accordingly. But with increasing membrane age, the salt
rejection rate of the membrane elements decreases, i.e. the TDS and concentration of
the saline components in the product water rise. For a two-pass seawater desalination
plant, this increases the capacity needed for the post-desalination stage and thus also
the plant’s capital costs and power demand. The power demand of the first pass
likewise increases because, the longer the AMLT, the greater is the operating
pressure needed to compensate for the permeability decline of the membrane
elements there installed.
5.3 Fouling and Scaling in RO Systems 485

A membrane element replacement schedule as shown in Table 5.16 is a planning


tool with unvarying and therefore idealized replacement rates as well as simplified
membrane management procedures. For the actual operation of a membrane desali-
nation plant, naturally the prevailing and also frequently varying seawater quality,
the resulting fouling potential, and other plant-specific influences on membrane
ageing dictate the sequence and extent of membrane replacement. Also, of course,
the membrane replacement procedure itself as well as how the elements in the
membrane modules are changed out influence the membranes’ age structure within
an RO array. Usually, not all of the module’s elements are changed out at once.
Instead, because of their shorter service life, the front elements are replaced more
frequently. For membrane modules whose elements are all identical, often the
downstream ones that are not subjected to such a high TDS concentration are shifted
to the front with the new elements then installed at the rear of the module. Because of
the differing types of element for modules in a hybrid configuration, it is not possible
to replace the membranes in this way. In this case, during membrane replacement the
membrane module is fitted with new elements at both the feed and discharge ends.
Of course, such membrane management requires that records be kept of each
membrane element. This involves identifying their positions the first time the plant is
loaded and noting changes in location and membrane lifetime in the course of
membrane replacements. With large-scale plants for which detailed tracking of the
location and age of each membrane element is no longer practical by manual logging
because of their large number, membrane management is tied into the overall plant
management system as a separate database.
The selection of the membrane age or the membrane replacement rate associated
with it is, in addition to the options for plant optimization at the membrane module
and array level described above, another possibility to influence the cost structure
and performance of a reverse osmosis system, this time at system level.
If these options are exploited, though, improvements of RO system performance,
salt rejection, and energy consumption are often associated with an increase in
equipment and/or operational costs. Weather and to which extent the advanced
measures for system optimization as described find application should therefore
always be decided on the basis of the cost benefits of the optimization measures
itself and a possible increase in the costs for additional expenditure in terms of
equipment and in operational costs. These cost comparisons should be made over the
entire life cycle of the desalination plant (live cycle cost analysis).

5.3 Fouling and Scaling in RO Systems

Suspended and colloidal pollutants in the feed line to an RO system as well as


particulates that are formed together with the build-up of saline concentration as
desalination proceeds result in deposition of coatings on membrane surfaces and
accretion of deposits in the flow channels of the membrane elements.
When investigating the causes of the formation of such coatings, a differentiation
is made between fouling and scaling. Understood as fouling are those deposits that
486 5 Reverse Osmosis Membrane System: Core Process of SWRO

form due to the ingress into the membrane elements of particulates and
macromolecules that are already present in the feedwater. Scaling arises when the
solubility of ionogenic compounds that are only slightly soluble—termed
“scalants”—is exceeded due to the solution becoming more concentrated during
desalination so that particulate precipitants of low solubility are formed.
The formation of such coatings on the membrane surface reduces membrane
permeability, but also affects their salt rejection, which may be appreciably impaired
depending on the extent of fouling or scaling, as described in detail below. Deposits
in the flow channels of spiral-wound membrane elements in particular as well as at
and in their spacers result in increased pressure loss within these elements.

5.3.1 Types of Fouling and Scaling with Their Influencing


Parameters

Deposits characterized as fouling build up in a membrane module or an array mostly


at those membrane elements that are nearest to the RO unit’s feed line, i.e. those at
the front end. Here, the operating conditions for extensive deposition of particulate
pollutants are most acute (see Sect. 5.3.2.1). But this does not exclude the possibility
of pollutants, present in particular as stable colloids, being swept further into the
downstream membrane elements and being deposited on these. Biological fouling
may extend over all membrane elements and the entire array of a membrane system.
The formation of coatings by scaling occurs where the concentration of ionogenic
constituents of scalants is highest, this being in the rearmost membrane elements.
Especially fouling but also scaling are complex processes in which greatly
dissimilar inorganic, organic, and also biologically active substances in the form
of particles, colloids, and macromolecules play a part. Alongside the differing nature
of foulants and scalants, there are other influencing factors:

• differing particle sizes and particle size ranges of foulants and scalants
• the particulate or colloidal structure of foulants and, if colloids are present, their
stability
• the relative electrical charges between the particles, the particle types, and the
membranes, how these influence the adhesion and repulsion of particles at the
membranes and between each other, and thus, as a result of the latter, the changes
in particle size due to their agglomeration
• other physical interactions between particulate and colloidal particles such as
mutual influencing by van der Waals forces as well as hydrophobic, hydrophilic,
and steric effects
• the physical structure of the membranes regarding surface characteristics and
roughness
• the hydraulic conditions in the flow channel and at the membrane elements’
surfaces
• the influence of salt concentration, i.e. the ionic strength
5.3 Fouling and Scaling in RO Systems 487

• the presence of organic compounds in the form of macromolecules or of


compounds from which macromolecules can form with the resulting deposition
of very dense membrane coverings
• nutrients swept into the membrane system that promote biological growth and
thus biofouling
• mutual influencing of fouling and scaling, for instance, build-up of excessive
concentration levels due to increased fouling and initiation or boosting of scaling
in this region.

5.3.2 Mechanism of Particle Transport, Deposition of Particles


on Membranes, and Properties of the Deposits

The primary process for both fouling and scaling is the deposition of particles on a
membrane surface. Despite the many factors that affect the nature and extent of
coating formation on the membranes in practice, simplified investigations and
modelling of the transportation of non-interactive particles to membrane surfaces
and their accretion there can give an idea of and provide an insight into the
parameters that influence the physical conditions when particles are deposited on
permeable membranes.

5.3.2.1 Particle Transport and Deposition


Like for ionogenic substances, for particle deposition from a flow channel onto a
membrane surface, the dependencies on concentration polarization are as described
in Sect. 5.1.5.1.3. However, particles are completely retained at the membrane
surface. There then results a concentration polarization factor of particle deposition
βp that is equal to the ratio of particle concentration at the membrane wall cp,W to the
particle’s concentration in the bulk suspension cp,B, and this depends on the product
flux JW and a factor kDp, this being the mass transfer coefficient of the particulate
substance concerned (see Eq. 5.128).

cp,W JW
¼ ekDp ¼ βp : ð5:128Þ
cp,B

cp,W ¼ particle concentration at membrane wall [kg/m3] [g/l]


cp,B ¼ particle bulk (concentrate) concentration [kg/m3] [g/l]
JW ¼ product flux [m3/m2h], [m3/m2s]
kDp ¼ mass transfer coefficient of particulate [m/h], [m/s]
βp ¼ concentration polarization factor of particulate deposition []
In turn, kDp results from the diffusion coefficient of the particle Dp, which governs
its transport to the membrane and its return transport back into the bulk suspension as
well as the thickness δp of the coating forming on the membrane (see Eq. 5.128a).
488 5 Reverse Osmosis Membrane System: Core Process of SWRO

Dp
kDp ¼ : ð5:128aÞ
δp

Dp ¼ shear-induced diffusion coefficient of particle [m2/s] [m2/h]


δp ¼ thickness of particle layer/cake [m]
Depending on the size of the particle, there are various transport mechanisms that
could model its return transport from the wall. For particles of diameter greater than
0.2 μm and large colloids, a shear-induced mechanism is assumed. In this case, Dp
results from Eq. (5.128b). Dp then depends on the effective flow velocity υF,C,eff
within the membrane element’s flow channel, the square of the radius of the particle
ap in the bulk suspension, and the coating thickness δp [77].

0:2  υF,C,eff  a2p


Dp ¼ : ð5:128bÞ
δp

υF,C,eff ¼ effective feed to concentrate velocity [m/s]


ap ¼ particle radius [m]
For a flow channel fitted with a spacer, under consideration of the channel
geometry as well as the geometry and characteristics of the spacer, the effective
velocity υF,C,eff in the channel is calculated with Eqs. (5.53a), (5.59c), (5.59d),
(5.59e), and (5.59f).
The concentration of the particles at the membrane wall is calculated from the
particle concentration in the feed line to the membrane element cp,F, the concentra-
tion factor CFP, and βp, as shown in Eq. (5.128c).

cp,W ¼ cp,B  βp ¼ cp,F  CFp  βp ð5:128cÞ

cp,F ¼ particle concentration in feed [kg/m3] [g/l]


CFp ¼ concentration factor for particles (Rp ¼ 100%) []
To determine the general principles for particle transport and deposition at a
membrane surface, investigations are conducted with synthetic suspensions of
organic and inorganic particles of a defined particle size, in particular with
microfiltration and ultrafiltration membranes under both dead-end and crossflow
conditions [78]. The crossflow mode corresponds to the conditions within a contin-
uously operated RO system. From such trials in the crossflow mode, there results the
definition of a critical flux Jcrit, a parameter that represents the value of the membrane
permeate flux below which there is no or only slight deposition of particles onto its
surface [79]. If for a certain particle type and particle size this value for the critical
flux is exceeded, scaling of the membrane surface intensifies.
To determine the critical flux Jcrit, i.e. the membrane flux at which deposition on
the membrane surface just commences, various laboratory test methods find appli-
cation. For MF and UF high-permeability membranes, the most common is to plot a
graph of trans-membrane pressure (TMP) against membrane flux. The curve shows a
change when particle deposition commences. This is termed TMP profiling. Another
test method suitable for membranes that, compared to the throughflow resistance of
5.3 Fouling and Scaling in RO Systems 489

the coatings that form, exhibit their own significantly greater resistance,
i.e. nanofiltration and RO membranes, is direct observation of particle behaviour
through a transparent membrane, referred to as direct observation through membrane
(DOTM). Other methods employ ultrasound, optical lasers, or tracer substances to
detect the onset of deposition [77, 80].
The critical flux Jcrit is calculated with Eq. (5.129), which is a modification of
Eq. (5.128), but with a shear-induced mass coefficient kS instead of the mass
coefficient kDp. By applying a modified shear-induced diffusion model, kS is calcu-
lated for particles of diameter 5–12 μm with Eq. (5.129a) and for colloids and
particles of less than 5 μm diameter with Eq. (5.129b). The shear rate γ at the
membrane wall of a rectangular flow channel is determined from the mean effective
velocity within the channel υeff and its hydraulic diameter dH (Eq. 5.129c). For the
above ranges of particle diameter, this calculation model, i.e. Equation (5.129),
(5.129a), (5.129b), and (5.129c), shows good agreement with the values of critical
flux as measured in lab trials [78].

cp,W
J crit ¼ ks  ln ¼ ks  ln βp : ð5:129Þ
cp,B
!13
a4p
ks ¼ α  γ  W : ð5:129aÞ
l

!13
γ  a2p
ks ¼ 0:807  : ð5:129bÞ
l

υF,C,eff
γ ¼6 : ð5:129cÞ
dH
Jcrit ¼ critical flux [m3/m2, h], [m3/m2, s], [m/h], [m/s]
ks ¼ shear-induced diffusivity mass coefficient of deposition [m/h], [m/s]
l ¼ length of flow channel [m]
dH ¼ hydraulic diameter of flow channel [m]
γ ¼ shear rate at wall of flow channel [s1]
For flow channels fitted with spacer fabrics, their hydraulic diameter dH is
calculated with Eqs. (5.59f), (5.59g), and (5.59h) or also with Eqs. (5.59e), (5.59i),
and (5.59j) depending on the spacer’s characteristics as well as its geometry and that
of the channel.
Due to the spacer’s influence on the hydraulic conditions in a flow channel, there
is increases in effective flow velocity, turbulence, and shear rate at the membrane
wall, also the value of the critical flux is raised compared with what it would be in an
unobstructed flow channel.
490 5 Reverse Osmosis Membrane System: Core Process of SWRO

Equation (5.129d) shows how the geometry and orientation of fabric spacers
influence the critical flux in a spiral-wound module for organic particles of diameter
6 μm [81].

J crit ¼ α þ β  Re : ð5:129dÞ

Orientation of spacer α β
45 +58.80 0.3702
90 37.63 0.5698
0 +82.74 0.3988

During transportation and deposition of particles on a membrane, the mass of


particulate material mp,d,ΔτD deposited per square meter of membrane surface results
from the particle concentration cP,F in the feed to the membrane element and the net
flux ΔJP, which is the difference between the flux JW and the critical flux Jcrit of the
particulate substance, as well as the time period ΔτD over which deposition takes
place (Eq. 5.130). For this, in actual practical operation of membrane systems, ΔτD is
the time interval between chemical cleaning of the membranes. The higher the net
flux ΔJW, i.e. the more the membrane flux JW exceeds the particles’ critical flux Jcrit,
the more material will be deposited on the membrane surface. If the net flux ΔJW is
negative, material will be ablated from the membrane surface.
Further, the specific mass of deposited material mp,d,ΔτD has to be multiplied by a
fractional deposition coefficient Ω since during a membrane plant’s crossflow
operation, but depending on hydraulic conditions in the flow channel, part of the
deposited material is swept back into the flow and leaves the membrane element. Ω
therefore depends on the flow parameters in the membrane element’s flow channels.
This coefficient is unity if all particles are deposited and is less than unity if only
some of the particles are deposited. Ω becomes negative if all deposits on the
membrane are again swept off [82].

mp,d,ΔτD ¼ cp,F  ðJ W  J crit Þ  Ω  ΔτD : ð5:130Þ

Ω ¼ fractional deposition coefficient []


mp, d, Δτ ¼ mass of deposited particulate matter per membrane area [kg/m2],
[g/m2]
mp,d,ΔτD ¼ mass of deposited particulate matter per membrane area [kg/m2],
[g/m2]
In accordance with this model for the transport and deposition of particles on a
membrane surface, the extent of deposition depends on their concentrations in the
membrane feed and also by how far at a certain point of the membrane surface a
critical flux has been exceeded for a specific particulate substance. Thereby the value
of this critical flux is fixed by:
5.3 Fouling and Scaling in RO Systems 491

• the particle’s radius ap, to either its second or fourth power


• the ratio of particle concentration at the membrane surface to that in the mem-
brane feed, that is the concentration polarization factor for particle deposition βp
• the shear rate at the membrane surface γ, that is the flow conditions prevailing in
the flow channel.

Smaller particles are deposited much more easily than is the case for larger
particles. Deposition is favoured at points of the membrane surface with high flux
and low shear rate, i.e. where the effective velocity or also crossflow velocity υF,C,eff
is low.
This is also the explanation for when, at an RO membrane module with more than
one element connected in series on the concentrate side, particulate deposition is
greater at the frontmost elements as here the specific flux JW is at a maximum, which
means its difference from the critical flux, i.e. the net flux ΔJW, is greatest. The flux
then drops in the downstream elements and thus also the potential for fouling (see
Figs. 5.37 and 5.43b).
This simplified model also explains the reduction of the degree of fouling in the
front elements, if perhaps due to a hybrid configuration of the membrane modules or
their multi-stage configuration in an array, the specific flux is less there. Also, if the
recovery rate of a membrane module or of an array drops in the front modules, the
specific flux as well as the average specific flux of the array decreases, so reducing
the fouling potential. The cleansing effect of low-pressure flushing of membrane
elements at a correspondingly low flux and high crossflow, that is with increased
Reynolds number Re and shear rate γ, at the membrane surface can likewise be
derived from the critical flux model.
For transfer of the critical flux model to RO systems and then specifically to
organic and inorganic membrane fouling processes, a number of lab trials have been
conducted, as described above. Used for these trials employing the DOTM detection
process with transparent MF membranes were synthetic suspensions of latex
particles of size from 3 to 11.9 μm and of silicate particles as well as of yeast cells
and algae or suspensions of selected bacterial species.
For trials with membranes fitted with spacers, depending on the crossflow
velocity, i.e. the Re number, in the test cells, the suspensions’ particle concentration,
the particle size, and the spacer orientation, during various trial runs with latex
particles of size 6.4 μm, values for the critical flux of between 10 and 40 l/m2, h were
determined. For suspensions of bacteria and algae, from trial runs with comparable
crossflow velocities, values for the critical flux have been published with one from
less than 10 to 40 l/m2, h, but also up to over 400 l/m2, h [83, 84].
During test runs with colloidal silicate particles of diameter 0.02 μm in flow
channels fitted with spacers and employing an NaCl tracer method, critical flux
values of between 10 l/m2, h and 32 l/m2, h were measured [85]. Other investigations
revealed a critical flux of around 20–25 l/m2, h for iron hydroxide particles [80].
In another literature source, for particles of diameter between 0.2 and 1.0 μm,
critical flux values of from 1 to 14 l/m2, h were quoted, while for other tests for the
same particle size range, the figures were 1 to 18 l/m2, h [77, 86].
492 5 Reverse Osmosis Membrane System: Core Process of SWRO

The bandwidths of these quoted values reveal that lab trials with suspensions of
synthetic mono-particles yield quantitative values that are only of very limited
application for the design of RO plants. In any case, fouling with its large number
of influencing factors is too complex.
Thus, these trials also show that the critical flux for organic substances in
particular depends greatly on their zeta potential and, for both organically inactive
and biologically active substances, the factors described above for the formation of
macromolecules and agglomerations as well as of physical interactions between the
particles themselves and with the membranes significantly influence the deposition
characteristics [87]. Further, alongside deposition on the membrane surface for
biological fouling in particular, adhesion of biological and inorganic materials to
the spacers of the membrane elements greatly influences membrane behaviour [88].
For seawater desalination in practice, the fouling agents with which the mem-
brane elements come into contact are made up of a mixture of suspended and
colloidal substances of diverse particle sizes and differing characteristics. Such a
mixture of substances may consist of inorganic and organic constituents as they
occur in seawater as well as coagulants and flocculants left over from its
pretreatment, plus biologically active material. The content of organic and biological
substances relative to that of inorganic material depends greatly on climatic and
weather-related seawater conditions as well as on the result of pretreatment under
these conditions. If scaling occurs, the substances that cause this are to a large part
inorganic compounds of the scalants alkaline earth sulphates and calcium carbonate.
Thus, how the various particle types influence each other and their behaviour when
being deposited onto membranes and spacers can differ, and also change,
accordingly.
Also to be taken into account is that the operating conditions regarding specific
flux, flow patterns, particle concentration and size, and TDS in the frontmost
elements of a membrane module continuously vary, thus modifying these elements’
critical flux.
This means that values for the critical flux applied when designing the desalina-
tion system and that may also be used as guide values for their operation will have to
be determined for conditions that approximate as closely as possible membrane
desalination operation in actual practice. This can be done by pilot trials with
membrane elements or membrane modules to which raw seawater is admitted or
by taking measurements at existing plants.
But when applying the critical flux concept for dimensioning seawater desalina-
tion plants in practice, it must also be borne in mind that not only the membrane
manufacturers’ stipulations and limit values have to be complied with for membrane
design (see Table 5.12), but also commercial aspects, i.e. CAPEX and OPEX, have
to be considered.
The specific flux that results for the frontmost elements of a seawater desalination
system under the selected design values for the product recovery rate and the average
system specific flux as well as seawater TDS normally lies within a range of from
15 to 30 l/m2, h. If when applying the critical flux concept, it should prove necessary
to operate an RO system at a specific flux for the frontmost elements which is
5.3 Fouling and Scaling in RO Systems 493

substantially reduced from the results of the membrane manufacturer’s design


software; the consequence for the RO system is that it will have to be equipped
with a larger membrane area compared to the manufacturer’s dimensioning and
potentially there could be a conflict with the manufacturer’s design values. In such a
case it will be necessary to find a compromise between the techno-economic aspects
of the design and reduction of the fouling potential. This could be done by dimen-
sioning the RO system for a sustainable flux at the frontmost elements that would
exceed the critical flux, but nevertheless still permit minimization of fouling.

5.3.2.2 Properties of the Deposits


Deposits on the membrane surface and within the flow channels of RO membrane
elements affect their product flux and salt rejection. Because of the reduction of
membrane permeability due to scaling and increased pressure loss as a consequence
of throughflow impairment, also the feed pressure needed at the elements has to be
raised. The extent to which and how these influences on system performance
manifest themselves depend on the nature of the deposits and the magnitude of the
resulting fouling.

5.3.2.2.1 Resistance of Membrane Deposits


According to the resistance-in-series model, the total resistance to flow rT through a
fouled RO membrane is the sum of the flow resistance of the membrane rM and the
resistance of the coating that has formed on the membrane surface r C,ΔτD consisting
of particulate material, i.e. the particle cake, as shown in Eq. (5.131). The flow
resistance of the membrane rM is the reciprocal of its permeability AM, see
Eq. (5.131a). By combining Eqs. (5.131) and (5.131a), and the basic equation for
the flux JW of an RO membrane, there results the relationship shown in Eq. (5.131b)
for the membrane flow resistance rM as a function of the pressure differential ΔpM
and osmotic pressure difference Δπ M at the membrane together with the membrane
flux JW.

ΔpT,ΔτD  Δπ M ΔpT,ΔτD  Δπ M
JW ¼ ¼ : ð5:131Þ
μF  r T μF  ðr M þ r C,ΔτD Þ
1
rM ¼ : ð5:131aÞ
AM  μF
1 ΔpM  Δπ M
rM ¼ ¼ : ð5:131bÞ
J W μF J W  μF
ΔpM Δπ M

rT ¼ total resistance of fouled membrane [m1]


rM ¼ membrane resistance [m1]
r C,ΔτD ¼ cake resistance at deposition time Δτ [m1]
JW ¼ membrane flux [m3/m2s] [m/s]
AM ¼ water permeability coefficient of membrane [m2s/kg]
μF ¼ dynamic viscosity of membrane feed [kg/sm] [Pas] (Ns/m2]
494 5 Reverse Osmosis Membrane System: Core Process of SWRO

Then, as shown in Eq. (5.131c), the membrane flux JW will be determined by the
total pressure differential ΔpT,ΔτD composed of the pressure differential ΔpM as
needed by the “clean” membrane to overcome the osmotic pressure and its flow
resistance together with a pressure component ΔpC,ΔτD to compensate for the flow
resistance of the fouling cake (Eqs. 5.131d, 5.131e, and 5.131f). ΔpC,ΔτD in turn
depends on the time over which the deposit has formed ΔτD, whose maximum value
is the time interval between chemical membrane cleaning operations. In step with the
flow resistance of the fouling cake r C,ΔτD that increases with deposition time ΔτD,
also the total pressure differential ΔpT,ΔτD needed for the membrane element to
maintain the membrane flux JW rises over time (Eq. 5.131e).

ΔpT,ΔτD  Δπ M
J W ¼ Δp : ð5:131cÞ
M Δπ M
JW þ μF  r C,ΔτD

ΔpT,ΔτD ¼ ΔpM þ μF  J W  r C,ΔτD : ð5:131dÞ

ΔpC,ΔτD ¼ μF  J W  r C,ΔτD : ð5:131eÞ

ΔpT,ΔτD ¼ ΔpM þ ΔpC,ΔτD : ð5:131fÞ

ΔpT,ΔτD ¼ total membrane pressure differential at deposition time ΔτD


[Pa] [kg/ms2] [N/m2]
ΔpM ¼ pressure difference of clean membrane [Pa] [kg/ms2] [N/m2]
Δπ M ¼ osmotic pressure difference at clean membrane wall [Pa] [kg/ms2]
[N/m2]
ΔpC,ΔτD ¼ pressure differential of deposit (cake) at deposition time ΔτD
[Pa] [kg/ms2] [N/m2]
The resistance to flow through the fouling cake r C,ΔτD is calculated from the
specific deposition quantity mp,d,ΔτD and the specific flow resistance of the cake αC
(Eq. 5.132). αC is a parameter that is specific for a particle layer and it can be
estimated for uniform spherical particles and non-compressible coatings on the basis
of the Carman-Kozeny relationship from the porosity ε of the deposit together with
the density ρp and the diameter dp of the particles that it consists of Eq. (5.132a).

r C,ΔτD ¼ mp,d,ΔτD  αC : ð5:132Þ

180  ð1  εÞ
αC ¼ : ð5:132aÞ
ρp  d2p  ε3

αC ¼ specific resistance of cake [m/kg]


ε ¼ porosity of deposition layer []
dp ¼ particle diameter [m]
ρp ¼ density of particle [kg/m3]
Equation (5.132b) that is derived from Eqs. (5.132) and (5.132a) shows that the
flow resistance r C,ΔτD is a function of the specific deposition quantity mp,d,ΔτD and the
5.3 Fouling and Scaling in RO Systems 495

coating’s throughflow characteristics. The specific deposition quantity may also be


characterized by the cake thickness δC,ΔτD , its porosity ε, and the density ρp of the
particles that make it up (Eq. 5.132c). From this equation together with Eqs. (5.132)
and (5.132d) is derived that shows the dependency of the flow resistance rC, Δτ on the
cake thickness δC, Δτ, its porosity ε, and the particle diameter dp.

180  ð1  εÞ
r C,ΔτD ¼ mp,d,ΔτD  : ð5:132bÞ
ρp  d2p  ε3

mp,d,ΔτD ¼ δC,ΔτD  ð1  εÞ  ρp : ð5:132cÞ

180  ð1  εÞ2
r C,ΔτD ¼ δC,ΔτD  ð5:132dÞ
d2p  ε3

δC,ΔτD ¼ thickness of cake at deposition time Δτ [m]


By combining Eq. (5.130) with Eq. (5.132), a relationship (Eq. 5.133) is obtained
for r C,ΔτD that also shows for this parameter the influence of the critical flux Jcrit, the
deposition time ΔτD, the concentration of the particles cp,F in the feed line to the
membrane, the fractional deposition coefficient Ω, and the specific flow resistance
αC. Equation (5.132a) together with Eq. (5.133) results in Eq. (5.133a) that addition-
ally reveals the influence of porosity ε as well as the density ρp and diameter dp of the
cake particles.

r C,ΔτD ¼ ðJ W  J crit Þ  ΔτD  Ω  cp,F  αC : ð5:133Þ

180  ð1  εÞ
r C,ΔτD ¼ ðJ W  J crit Þ  ΔτD  Ω  cp,F  : ð5:133aÞ
ρp  d2p  ε3

180  ð1  εÞ
ΔpC,ΔτD ¼ μF  J W  ðJ W  J crit Þ  ΔτD  Ω  cp,F  : ð5:134Þ
ρp  d2p  ε3

Like other models for determining membrane fouling characteristics, the model
described above for calculating the reduction in membrane permeability due to its
coating with a fouling cake relies on simplifying assumptions such as identical
particle sizes and a uniform particle structure within the layer together with its
incompressibility. Also, interactions between the particles and of these with the
membrane as well as other parameters that influence fouling are not considered (see
Sect. 5.3.1). Additionally, parameters are needed for such modelling whose values
for calculating a membrane system in practice can only be determined with difficulty
and then imprecisely or which vary depending on the conditions under which the
membranes operate, for instance porosity ε, density ρp, and diameter dp. Other
calculation parameters must be determined empirically by trials, such as critical
flux Jcrit and the fractional deposition coefficient Ø). Nevertheless, such idealized
models give an idea of how a range of physical parameters impact membrane
496 5 Reverse Osmosis Membrane System: Core Process of SWRO

behaviour. Thus, Eq. (5.134) shows how the difference between the flux JW and the
critical flux Jcrit, i.e. the net flux ΔJW; the deposition time ΔτD; the extent of
deposition as characterized by the fractional deposition coefficient Ø; the particle
concentration cp,F in the membrane feed; and the characteristics of the coating—
porosity, particle density and particle size—individually and in concert impact the
pressure loss of the fouling cake ΔpC,ΔτD and thus also the membrane feed pressure
ΔpT,ΔτD (Eq. 5.131f) that is needed to maintain the specified flux.
This reveals that in particular a small particle diameter resulting in a correspond-
ingly low porosity in the cake raises its flow resistance. This is equally the case if the
porosity presented by the particles is diminished by the deposition of organic
macromolecules or gelatinous substances in the interstices and filling of these, as
is often the case with organic and biological fouling. Large particles with a higher
density lessen the specific flow resistance (Eq. 5.132b) and thus also the pressure
loss. Obviously, longer deposition times will exacerbate cake formation and also
result in an increase of operation far above the critical flux.
When considering the fouling behaviour of RO membranes, in addition to the
reduction of membrane permeability as shown by this cake model is the influence of
the accruing coating on the ionogenic concentration of the solution components at
the membrane wall with the associated rise in concentration polarization, i.e. the
emergence of a so-called cake-enhanced osmotic pressure (CEOP).

5.3.2.2.2 Cake-Enhanced Osmotic Pressure


The formation of a coating of porous solids on the surface of a separation membrane
swept with a solution influences the concentration polarization of its ionogenic
constituents at the membrane surface (see Sect. 5.1.5.1.3). The deposition coating
reduces the degree of reverse diffusion of the ionogenic substances back into the
bulk solution with the consequence that more of these are retained at the membrane
surface. This process is also termed cake-enhanced osmotic pressure (CEOP) or,
more generally, fouling-enhanced osmotic pressure (FEOP) [89, 90].
The extent to which the diffusion of the components, or their diffusion coefficient
DSi,cake, influences the return transport back into the bulk solution depends on the
structure, i.e. the porosity ε, of the deposition coating (see Eq. 5.135).
 
ε
DSi,cake ¼ DSi  : ð5:135Þ
1  ln ε2
DSi ¼ diffusion coefficient of component i at clean membrane [m2/h], [m2/s]
DSi,cake ¼ reduced diffusion coefficient of component i in cake layer [m2/s], [m2/
h]
ε ¼ cake voidage, porosity []
The ratio of the concentration of component i of the solution at the membrane
wall ci,M,ΔτD and that in the bulk solution ci,B as well as its concentration ci,p,CEOP,ΔτD
in the product of the desalination membrane is fixed by the value of the polarization
modulus M P,CEOP,ΔτD in accordance with Eq. (5.136).
5.3 Fouling and Scaling in RO Systems 497

ci,M,ΔτD  ci,p,CEOP,ΔτD JW

¼ ekCP,Δτd ¼ M P,CEOP,ΔτD : ð5:136Þ


ci,B  ci,p,CEOP,ΔτD

M P,CEOP,ΔτD ¼ polarization modulus of fouled membrane due to CEOP at


deposition time ΔτD []
kCP,ΔτD ¼ concentration polarization mass transfer coefficient of fouled mem-
brane at deposition time ΔτD [m/h], [m/s]
ci,M,ΔτD ¼ concentration of component i at membrane wall of fouled membrane at
deposition time ΔτD [kg/m3] [g/l]
ci,p,CEOP,ΔτD ¼ concentration of component i in product of fouled membrane at
deposition time ΔτD [kg/m3] [g/l]
To calculate M P,CEOP,ΔτD , it is necessary to know the mass transfer coefficient
kCP,ΔτD that is determined from the mass transfer coefficient of the “clean” membrane
kCP,clean and that of its surface coating k CP,cake,ΔτD . k CP,ΔτD is a combination of these
two parameters as shown in Eqs. (5.136a) and (5.136b) [83].

1 1 1
¼ þ : ð5:136aÞ
kCP,ΔτD kCP,clean k CP,cake,ΔτD
kCP,cake,ΔτD
k CP,ΔτD ¼ kCP,cake,ΔτD
: ð5:136bÞ
1þ k CP,clean

kCP,clean ¼ concentration polarization mass transfer coefficient of clean membrane


[m/h], [m/s]
kCP,cake,ΔτD ¼ concentration polarization mass transfer coefficient of cake at
deposition time ΔτD [m/h], [m/s]
The mass transfer coefficient of the coating concentration polarization kCP,cake,ΔτD
results from the diffusion coefficient DSi,cake of the cake and its thickness δC,ΔτD in
accordance with Eq. (5.136c). If Eqs. (5.136c) and (5.132c) are combined to
form Eq. (5.136d), the general dependence of kCP,cake,ΔτD on the diffusion coefficient
DSi,cake, the coating porosity ε, the density ρp of the particles it contains, and the
specific quantity of deposits on the membrane mp,d,ΔτD becomes apparent.
 
DSi,cake DSi  1 lnε ε2
k CP,cake,ΔτD ¼ ¼ : ð5:136cÞ
δC,ΔτD δC,ΔτD
 
DSi  1 lnε ε2  ð1  εÞ  ρp
kCP,cake,ΔτD ¼ : ð5:136dÞ
mp,d,ΔτD

kCP,cake,ΔτD ¼ concentration polarization mass transfer coefficient of cake at


deposition time ΔτD [m/h], [m/s]
δC,ΔτD ¼ thickness of cake at deposition time ΔτD [m]
ρp ¼ density of particles [kg/m3]
498 5 Reverse Osmosis Membrane System: Core Process of SWRO

The mass transfer coefficient kCP,clean of the “clean” membrane can be calculated
as shown in the basic RO equations under Sect. 5.1.5.1.3, these being Eqs. (5.33),
(5.33a), and (5.33b).
It is also possible to derive from Eq. (5.135d) together with Eq. (5.136b) the
following Eq. (5.136e) for the combined kCP,Δτ from kCP,clean of the “clean” mem-
brane and, depending on its physical parameters, the characteristics of the fouling
coat deposited on it.
" #1
mp,d,Δτ 1
kCP,Δτ ¼   þ : ð5:136eÞ
DSi  1 lnε ε2  ð1  εÞ  ρp kCP,clean

From kCP,clean and kCP, cake, Δτ, the combined mass transfer coefficient kCP,ΔτD can
be calculated with Eq. (5.136b) and then together with the membrane flux JW the
polarization modulus M P,CEOP,ΔτD of the fouled membrane is obtained in accordance
with Eq. (5.136).
To the extent that, during the fouling process, the specific deposition quantity
mp,d,ΔτD on the separation membrane increases, the diffusion coefficient DSi and thus
also the mass transfer coefficient kCP,ΔτD reduces, the value of the polarization
modulus Mp,clean then rises to the respective higher value M P,CEOP,ΔτD for the fouled
membrane. Thus, also the concentration of the solution component ci,M at the
membrane wall increases from the value ci,M,clean of the “clean” membrane to the
value ci,M,CEOP,Δτ for the fouled membrane.
Like for the flow resistance rC,Δτ that arises due to membrane fouling (Eq. 5.133a)
and the associated differential pressure ΔpC,Δτ (Eq. 5.134), the amount by which the
concentration on the membrane side increases through CEOP is influenced by the
fouling cake’s physical properties and in particular its porosity ε (Eqs. 5.136, 5.136b,
5.136c, 5.136d, and 5.136e). The lower this porosity, the more the diffusion coeffi-
cient DSi,cake in the cake is reduced (Eq. 5.135), so lessening the return transport of
the ionogenic components from the membrane wall into the bulk solution with the
result that the concentration ci,M,ΔτD at the membrane wall is raised.
If the desalination membranes are operated in the constant flux mode, the
magnitude of the CEOP effect generated by fouling may be determined from the
ratio of the polarization modulus of a fouled membrane M P,CEOP,ΔτD to that of a
“clean” membrane Mp,clean as the CEOP fouling factor F CEOP,ΔτD using Eq. (5.137).

M p,CEOP,ΔτD
¼ F CEOP,ΔτD : ð5:137Þ
M pclean

F CEOP,ΔτD ¼ CEOP induced fouling factor at deposition time ΔτD []


If a membrane’s intrinsic (i.e. referred to the membrane wall concentration ci,M)
salt passage SPiM or salt rejection RiM is known, its polarization modulus Mp may be
calculated directly from the concentration ci,F to which it is exposed, the concentra-
tion factor CF, and the measurable product concentration ci,p (Eqs. 5.138 and
5.138a).
5.3 Fouling and Scaling in RO Systems 499

ð1  SPiM Þ ci,p
Mp ¼  : ð5:138Þ
SPiM ci,F  CF  ci,p
RiM ci,p
Mp ¼  : ð5:138aÞ
ð1  RiM Þ ci,F  CF  ci,p

Mp ¼ polarization modulus of membrane []


SPiM ¼ membrane (intrinsic) salt passage factor of component i []
RiM ¼ membrane (intrinsic) salt rejection factor of component i []
A similar relationship is derived from Eq. (5.136) in order to calculate Mp from
the salt passage SPiM or the salt rejection RiM together with the concentration
polarization factor β (Eqs. 5.139 and 5.139a).

β  ð1  SPiM Þ
Mp ¼ : ð5:139Þ
1  β  SPiM
β  RiM
Mp ¼ : ð5:139aÞ
1  β  ð1  RiM Þ

The mass transfer coefficient kCP of a separation membrane and its associated
polarization modulus Mp may also be determined experimentally by a membrane
testing process, as described in [91, 92].
By combining Eq. (5.137) with Eq. (5.138) or rather Eq. (5.138a) for calculating
the respective polarization modulus of the fouled membrane M p,CEOP,ΔτD and that of
the “clean” membrane Mp, clean, from the particular product concentration
ci,p,CEOP,ΔτD and ci, p, clean, with Eq. (5.140) the CEOP fouling factor F CEOP,ΔτD that
characterizes the degree of the CEOP effect of the fouled membrane can be calcu-
lated. So that the two polarization modulus values can be compared, both
membranes have to have the same membrane flux JW while the other operating
parameters of temperature tF, TDS cF and recovery rate Y, or concentration factor CF
have to be normalized (see Sect. 5.3.3.2.1), which means they will then present the
same standard values.

ci,p,CEOP,ΔτD ðci,F CFci,p,clean Þ


F CEOP,ΔτD ¼  : ð5:140Þ
ci,p,clean  ci,F  CF  ci,p,CEOP,ΔτD

For seawater desalination, the value of the product concentration ci,p is negligible
compared to the value of the product from the feed concentration ci,F and the
concentration factor CF, i.e. the brine concentration ciB. Thus, Eq. (5.140) can be
simplified to Eq. (5.140a) which means that F CEOP,ΔτD is directly proportional to the
ratio of the product concentration ci,p,CEOP,ΔτD of the fouled membrane to that of the
“clean” membrane ci,p,clean.
500 5 Reverse Osmosis Membrane System: Core Process of SWRO

ci,p,CEOP,ΔτD
F CEOP,ΔτD ffi : ð5:140aÞ
ci,p,clean

Similarly, Eq. (5.137) can be combined with Eqs. (5.139) or (5.139a). In this case,
the CEOP factor F CEOP,ΔτD is then obtained as the ratio of the concentration
polarization factor βCEOP,ΔτD of the membrane with a fouling coating and the clean
membrane βclean (Eq. 5.140b).

βCEOP,ΔτD 1  βclean  SPiM


F CEOP,Δτ ¼  : ð5:140bÞ
βclean 1  βCEOP,ΔτD  SPiM

βCEOP,ΔτD ¼ concentration polarization factor of fouled membrane at deposition


time ΔτD []
βclean ¼ concentration polarization factor of clean membrane Δτ []
Due to the very low intrinsic salt passage through the RO seawater desalination
membrane, the product of βclean  SPiM or βCEOP,ΔτD  SPiM is negligible and so to a
1βclean SPiM
good approximation the quotient 1β SPiM can be set equal to one. There thus
CEOP,ΔτD

results the simplified Eq. (5.140c) for how FCEOP,Δτ depends on the concentration
polarization factors βCEOP,Δτ and βclean.

βCEOP,ΔτD
F CEOP,ΔτD ffi : ð5:140cÞ
βclean

If the concentration polarization factor βCEOP,ΔτD is known, with Eq. (5.141) the
respective component concentration ci,M,ΔτD at the wall of the fouled membrane at a
defined deposition time Δτ can be calculated from the feed concentration ci,F and the
concentration factor CF. Similarly, from the osmotic pressure in the feed line, it is
possible to calculate the osmotic pressure at the membrane wall at time ΔτD
(Eq. 5.141a).

ci,M,ΔτD ¼ ci,F  CF  βCEOP,ΔτD ¼ ci,F  CF  βclean  F CEOP,ΔτD : ð5:141Þ

π M,ΔτD ¼ π F  CF  βCEOP,ΔτD ¼ π F  CF  βclean  F CEOP,ΔτD : ð5:141aÞ

π M,ΔτD ¼ osmotic pressure at membrane wall of fouled membrane at deposition


time ΔτD [bar] [Pa] [kg/ms2] [N/m2]
π F ¼ osmotic pressure of membrane feed [bar] [Pa] [kg/ms2] [N/m2]
The CEOP effect raises the concentration ci,M,ΔτD of the ionogenic components at
the membrane wall and so influences both the membrane’s permeability and its salt
rejection. The value of the CEOP factor F CEOP,ΔτD is a measure of how the
concentration polarization of the fouled membrane is elevated after a specified
deposition time ΔτD due to CEOP and consequently and, when calculated for each
component of the solution to be desalinated, how their respective rejection rates and
also the permeability of the membrane are reduced.
5.3 Fouling and Scaling in RO Systems 501

The value of the increase Δπ CEOP,ΔτD of the osmotic pressure generated by the
fouling of the membrane and the CEOP effect is calculated from the difference of the
concentration polarization factors, βCEOP,ΔτD and βclean (Eq. 5.141b).
 
Δπ CEOP,ΔτD ¼ π F  CF  βCEOP,ΔτD  βclean
¼ π F  CF  βclean  ðF CEOP,ΔτD  1Þ: ð5:141bÞ

Δπ CEOP,ΔτD ¼ osmotic pressure increase due to the cake-enhanced osmotic pres-


sure CEOP at deposition time ΔτD [bar] [Pa] [kg/ms2] [N/m2]
If the CEOP effect is allowed for, the equation for the RO membrane in the
resistance-in-series model (Eq. 5.131) is then transformed to Eq. (5.142).

ΔpT,ΔτD  π M,ΔτD ΔpT,ΔτD  π M,ΔτD


J W,ΔτD ¼ ¼
μF  ðr M þ r C,ΔτD Þ ΔpM Δπ M,clean þ μF  r C,ΔτD
JW
ΔpT,ΔτD  βclean  F CEOP,ΔτD  π F  CF
¼ ΔpM βclean π F CF
: ð5:142Þ
JW þ μF  r C,ΔτD

J W,ΔτD ¼ membrane flux of fouled membrane at deposition time ΔτD [m3/m2s]


[m/s]
Both the raised flow resistance of a separation membrane ΔpT,ΔτD due to coating
formation and the resulting cake-enhanced osmotic pressure, CEOP, are influenced
by the velocity of the flow sweeping the membrane wall, but also very significantly
by the specific product flux JW to which the membrane is subjected. As shown by
Eq. (5.134) for ΔpT,ΔτD and the equations for calculating the CEOP effect
(in particular Eq. 5.136), however, the influence of product flux JW on the cake’s
flow resistance differs in intensity from its influence on the CEOP. According to
Eq. (5.134), the cake’s differential pressure ΔpC,ΔτD increases with the product of the
specific flux JW and the net flux ΔJW. But the polarization modulus M P,CEOP,ΔτD
induced by CEOP and mirroring this the concentration at the membrane wall
increase exponentially with the product flux JW (Eq. 5.136).
Thus, when designing a membrane system, the selected design values for the
specific flux JW,E for each membrane element and for the average specific flux JP,SØ,S
of the membrane system have to take into account the fouling potential that is still
present after precleaning the water to be desalinated (Table 5.12). If this residual
fouling potential is high, from the bandwidth of design parameters presented by the
membrane manufacturers, ones should be chosen that are more on the
conservative side.
When the RO membranes become fouled and especially for the high salinities
encountered in seawater desalination, it is the CEOP effect that impacts the
membranes’ salt rejection and permeability and, as a consequence when operating
at constant flux, also the operating pressure ΔpT,ΔτD needed for the separation
process as well as, but to a lesser extent, the flow resistance caused by the membrane
coating ΔpC,ΔτD [83, 85, 86, 93, 94].
502 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.3.3 Fouling

5.3.3.1 Fouling Potential of Raw and Pretreated Water: Testing Methods


To determine and characterize the fouling potential of a raw water to be desalinated,
testing methods have been developed in which flat sheet membrane filters are
charged with the relevant water at a certain pressure. For this test, the reduction of
filtrate flow is measured as the coating of its surface with fouling substances
increases. This is evaluated using the calculation algorithms specific for the testing
method and the result is then defined as the fouling potential.
The nature of the parameters and how they influence the filtrate flow in such a test
setup for this filtration method, which is also termed dead-end filtration, may be
depicted by the general filtration equation, this being the series resistance model in
accordance with Eq. (5.143).

dV f Δp  S f Δp  S f
¼ ¼  : ð5:143Þ
dτ f μFðtÞ,ðSÞ  r T μF ðtÞ,ðSÞ  r f þ r c

Vf ¼ filtrate volume [m3]


τf ¼ filtration time [s]
Δp ¼ pressure differential at filter [Pa] [kg/ms2] [N/m2]
Sf ¼ filter area [m2]
μF(t),(S) ¼ dynamic viscosity of filter feed at temperature t and salinity S [kg/sm]
[Ns/m2] [Pas]
rT ¼ total resistance of fouled filter [m1]
rτ ¼ total resistance of clean filter + filter cake [m1]
rf ¼ initial resistance of clean filter [m1]
rc ¼ resistance of filter cake [m1]
From Eq. (5.145), it can be seen that to obtain comparable values for how the
volumetric flow of the filtrate changes over the filtration time, for this testing method
it is necessary to specify the pressure at which the solution is admitted to the filtration
test rig, the filtration area, and the filter material’s flow resistance. This means that its
material properties like pore size and distribution, material type, and surface
characteristics (electrical charge and the degree to which it is hydrophilic or hydro-
phobic, etc.) have to be defined in an appropriate standard. Additionally, the
viscosity of the medium to be filtered and its dependence on the temperature and
TDS of the water under test significantly influence the filtration curve. If testing
conditions differ, it will have to be possible to convert their respective results to
generally accepted standard values in an algorithm included in such a standard.
For determining the fouling potential of a raw water, two standard test methods
based on such filtration tests find application, one being for the Silt Density Index
(SDI) and the other for the Modified Fouling Index (MFI). Both tests are conducted
using the same filtration setup (Fig. 5.45). They differ in how they record the data as
well as their algorithms for evaluating the measured data and for calculating the
fouling potential from the test results.
5.3 Fouling and Scaling in RO Systems 503

Fig. 5.45 Basic equipment


for SDI & MFI tests Feed

Pressure control valve PC

PI Pressure gauge

Filter holder

Membrane filter
Filtrate

The methodology for conducting the respective fouling test, that is:

• the basic equipment


• the stipulated test conditions, like water feed pressure, type and properties of the
filter material, etc.
• the procedure and conditions to be observed when conducting the test (for the SDI
test, how it is to be adapted to the fouling potential)
• how data are logged and the algorithms for evaluating the results of the
measurements

is laid down and described in the respective ASTM standards [95].


The SDI test is approved by all membrane manufacturers as the standard methodol-
ogy for characterizing fouling potential and has been used right from the start of the
application of membrane desalination technology as the industry standard for this
purpose as well as for monitoring operation of these plants.
So far, the Modified Fouling Index MFI has been applied mainly in scientific
works for characterizing and investigating various types of fouling and their causes
as well as for characterizing diverse fouling processes and in research in the field of
modelling fouling. However, it is now increasingly being used in technical
applications and for monitoring membrane systems for filtration and desalination.
Alongside these two fouling tests, R&D has come up with a number of other
fouling test methods and fouling algorithms with investigations into the characteri-
zation of fouling potential of waters and their suitability for predicting the fouling
behaviour of separation membranes, such as the Crossflow Sampler Fouling (CFS)
index, the Combined Fouling Index (CFI), etc. However, at present only the SDI and
504 5 Reverse Osmosis Membrane System: Core Process of SWRO

MFI testing methods have found industrial application in membrane technology and
membrane desalination for determining fouling potential.

5.3.3.1.1 Silt Density Index SDI


SDI Testing Method15
For measuring the Silt Density Index (SDI), the water whose fouling potential is to
be determined is passed through the testing rig shown in Fig. 5.45 at a constant
filtration pressure of 2.07 ∓ 0.07 bar (30 ∓ 1 psig). This is fitted with a standardized
flat sheet membrane filter that according to ASTM Standard D 4189 has a mean pore
size of 0.45 μm, a diameter of 47 mm, and a thickness of 115 to 180 μm. Addition-
ally specified in this ASTM standard is the filter material and its composition which
is cellulose nitrate mixed with cellulose acetate at a defined proportion, as well as a
range for the “pure water flow time” for a filtrate volume of 500 ml and thus the
permissible bandwidth of flow resistance of the membrane filter defined to be used
for the test.
Prior to starting the test with the water to be analysed, in line with the stipulations
of the ASTM standard, the time τRef is determined which is required by a filtrate
volume V1 of 500 ml of the “non-plugging reference water” to be filtered through the
test rig. Upon commencing the actual fouling test, first the time τ1 is measured for
which the same filtrate volume V1 like for the reference filtration of 500 ml of the
water under investigation passes through the filter. If the measured time τ1 is longer
by more than 1.1 times than the time τRef for the reference water, the test should be
repeated with a lesser filtrate volume V1, for example 250 or 100 ml, and then
conducted with this as test volume. After measuring τ1, the filtration test should be
continued and, after a defined time interval tf during which the filter is furthermore
charged with water with fouling substance, the time τ2 that is then required for
filtering the volume V2 is measured. V2 must then be the same as the filtrate volume
V1 as was chosen for determining τ1. This likewise applies for the temperature at
which the test is conducted and this should not vary by more than 1  C.
Depicted by means of a filtration curve that is typically generated in such a test for
how filtrate volume V changes with filtration time τ, Fig. 5.46 shows the sequence of
the time intervals τ1 and τ2 at which measurements are taken for the respective
preselected filtrate volume V1 (¼ V2) and the time interval τf from the start of the test
up to measurement of τ2. Calculated from the measured times τ1 and τ2 together with
the selected filtration time τf is the Silt Density Index SDIτ f , as shown by
Eq. (5.144).
 
%P30 1  ττ12  100
SDIτ f ¼ ¼ : ð5:144Þ
τf τf

15
ASTM D4189—Standard Test Method for Silt Density Index (SDI) in Water.
5.3 Fouling and Scaling in RO Systems 505

V/t [ml/min]

τf τ2
τ1

V1

V2
τ [min]

Fig. 5.46 Curve of filtrate volume flow vs. filtration time for SDI determination

SDIτ f ¼ silt density index at τf [%/min]


%P30 ¼ plugging rate [%]
τ1 ¼ time to collect accumulated Volume 1 [s]
τ2 ¼ time to collect accumulated Volume 2 after filtration time τf [s]
τf ¼ filtration time elapsed from start of SDI test up to start of collection
of Volume 2 [min]
Which filtration time τf is selected for the SDI test depends on the fouling
potential of the sample under investigation. For water with a low or average fouling
potential, for τf a time of 15 min is selected. If when conducting the test a value for %
P30 is measured that is greater than 75%, instead of 15 min for the filtration time a
shorter time of 10 or even 5 min should be set if the sample exhibits a higher fouling
potential, e.g. for surface-extracted untreated seawater, and then the test should be
repeated with this time.
When conducting an SDI test for seawater, appropriate corrosion-resistant
materials must be used for the test rig. This is because the products of corrosion of
such test rig materials or fouling of the equipment could indicate a higher fouling
potential of the water under investigation than is actually the case.
The manufacturers of RO membrane elements limit the fouling potential of the
water to be desalinated that is admitted to the elements by prescribing a maximum
value for SDI15. SDI15 is calculated with Eq. (5.144a). Depending on manufacturer
and element type, for the maximum permissible fouling potential the values for
SDI15 lie between 3 and 5. For permeate from a pre-desalination stage or following
filtration though MF or UF membranes, the maximum values for SDI15 for passage
through the downstream desalination stage should be lower (see Table 5.12).
506 5 Reverse Osmosis Membrane System: Core Process of SWRO

 
%P30 1  ττ12  100
SDI15 ¼ ¼ : ð5:144aÞ
15 15
SDI15 ¼ silt density index at τf ¼ 15 Min [%/min]

Drawbacks of the SDI Test and Proposed Changes


The SDI test does not require a high outlay for equipment and, with a straightforward
computing algorithm, the resulting measured data yield a parameter that
characterizes a water’s fouling potential under the test conditions. Although it has
been applied in practice for many decades as standard in membrane desalination
technology, the SDI test is flawed in quite a few aspects. Among its drawbacks are:

1. Experience has shown that the SDI is not directly proportional to the quantity of
particulate and colloidal pollutants present in the test water nor to its turbidity.
2. Due to it being passed through a 0.45 μm filter, not all constituents in a raw water
that contribute to fouling are included in the measurement. Colloidal constituents
in particular play a major role in membrane fouling, but their particle sizes are
significantly below the 0.45 μm limit.
3. Despite the mean pore diameter and material of the membrane filter being
specified in the ASTM standard, its flow resistance and particle deposition
characteristics can differ appreciably depending on filter manufacturer, even for
the same filtration batch [96, 97].
4. The flow of filtrate through a membrane is determined to a significant degree by
the viscosity of the test solution (Eq. 5.145). This viscosity in turn depends
greatly on temperature, but also, albeit to a lesser extent, on the TDS of the
water under investigation (see Sect. 3.2.2.2, Eqs. 3.41, 3.42, Fig. 3.16, and
Table 3.24). Consequently, changes in temperature and TDS influence the out-
come of any particular SDI test [97]. Especially when measuring the SDI of
seawater, not only the dependency of viscosity on temperature but also strictly
speaking on its TDS should be taken into account. The ASTM standard does not
include a correction algorithm that would permit normalization of the measured
SDI values with regard to a sample’s temperature and TDS.
5. According to ASTM Standard D 4189, it should only be used up to a turbidity
value of <1 NTU.
6. Differences in the pH of test solutions may also influence the SDI value if this
causes precipitation from the solution to occur [98].
7. With seawater that has been pretreated by ultrafiltration, an SDI value may be
measured that is higher than the sample’s actual fouling potential. The explana-
tion for this is that organic macromolecules in the filtered sample may be
deposited within the filter pores, thus restricting their passage and resulting in a
misleading and higher SDI measurement. This is also possible if micro air
bubbles that form during testing likewise partially block the pores.
5.3 Fouling and Scaling in RO Systems 507

Modifications as follows have been proposed for the SDI test that have also
already been put into practice:

• Silt Density Index based on a pluggage rate of the pores of 75% (SDI75%):
Upon exceeding the pluggage rate specified in the ASTM standard of 75%, the
SDI is referred not to a shortened filtration time τf of 5 or 10 min (designated
respectively as SDI5 or SDI10), but instead to a value for τf—designated τf,75%—
for which from the quotient ττ12 a plugging rate of 75% (%P75%) is calculated. The
Silt Density Index SDI75% then results from:
 
%P75% 1  ττ12  100
SDI75% ¼ ¼ : ð5:144bÞ
τ f 75% τ f 75%

SDI75% ¼ silt density index at 75% plugging rate [%/min]


τ75% ¼ filtration time at 75% plugging rate [min]
With the value so obtained for SDI75%, in particular for water with an average
or high fouling potential like raw seawater prior to precleaning and for quality
control of the precleaning process, etc., SDI values can be better compared with
each other [98].
• Volume-based Silt Density Index, SDIV:
For this parameter, the filtration interval between the start of the test and the
measurement of the volume V2 is not limited by time but instead by volume. This
means that the filtration time τ2 for a volume V2 is measured after a defined
standard volume VSt to be measured after starting the test (Eq. 5.144c). The reason
for this is that the extent to which the membrane is coated with fouling substances
up to the start of measurement of the volume V2 is fixed by the quantity of filtrate
that up to this time has passed through the filter and not so much by the filtration
time that has elapsed [96, 97].
 
%PV 1  ττ12  100
SDIV ¼ V St ¼ V St : ð5:144cÞ
Sf Sf

SDIV ¼ silt density index at VSt (volume-bases SDI) [%/m]

VSt ¼ filtration volume collected from start of SDI-test up to start of collection of


Volume 2 [m3]
Sf ¼ filter area [m2]

5.3.3.1.2 Modified Fouling Index MFI


MFI Basics
The calculation methodology for determining MFI is based on application of the
general filtration equation, i.e. the series resistance model of Eq. (5.145) and the
Carman-Kozeny relationship for cake filtration according to Eqs. (5.132a), (5.132b),
508 5 Reverse Osmosis Membrane System: Core Process of SWRO

τ/V [sec /l]

Filtration mechanism

Cake Cake filtration with


Blocking
filtration compression
filtration

tan α

MFI = slope = tan α

V [l]

Fig. 5.47 Curve of filtration time per volume vs. volume for MFI determination

and (5.132d) within the linear range of cake filtration of a filtration curve, with which
τ
the dependency of the quotient V ff of filtration time τf and the respective filtrate
volume Vf is depicted as a function of the filtrate volume Vf [82] (see Fig. 5.47).

dV f Δp
¼ Jf ¼  : ð5:145Þ
dτ f  S f μF ðtÞ,ðSÞ  r f þ r c

Jf ¼ filtrate flux during test [m3/m2s] [m/s] [l/m2s]


As Fig. 5.47 shows, initially the filtration curve traverses a region of blocking
filtration in which the filter’s membrane surface becomes fouled. As soon as the
surface becomes completely coated, the next phase of cake filtration commences for
which the relationship between filtrate volume and the filtration time is more or less
linear. This is followed by the phase of cake filtration accompanied by cake
compression for which the filtrate volume decreases in relation to the filtration time.
By combining Eq. (5.145) for the filtrate flux Jf during dead-end filtration with
Eq. (5.145a) for the flow resistance rc of the membrane’s coating, Eq. (5.145b)
results which, as well as the specific resistance of the cake αc, also allows for the
particle concentration cp,f in the medium being filtered.

Vf
rc ¼  αc  cp,F : ð5:145aÞ
Sf
5.3 Fouling and Scaling in RO Systems 509

αc ¼ Specific resistance of cake [m/kg]


cp,F ¼ Particle concentration in filter feed [kg/m3] [g/l]

dV f Δp
¼Jf ¼  : ð5:145bÞ
dτ f  S f V
μFðtÞ,ðSÞ  r f þ S ff  αc  cp,F

By integrating this relationship for a constant pressure (Δp ¼ const.) from τf ¼ 0


τ
to τf ¼ τf, a linear function for the dependency of V ff on Vf is obtained as shown by
Eq. (5.147c).

τf μF ðtÞ,ðSÞ  r f μF ðtÞ,ðSÞ  αc  cp,F


¼ þ  V f: ð5:145cÞ
Vf Δp  S f 2  Δp  S f 2

The value of the Modified Fouling Index MFI then corresponds to the value of the
coefficient of the filtrate volume Vf in Eq. (5.145c), which is the gradient of the linear
section of the graph during the cake filtration phase of Fig. 5.47 and is calculated
with Eq. (5.146). As shown by the Carman-Kozeny relationship (Eq. 5.132a), the
specific flow resistance of the cake αc depends on its porosity ε, the square of the
particle diameter dp, and the particle density ρp.

μF ðtÞ,ðSÞ  αc  cp,F
MFI ¼ : ð5:146Þ
2  Δp  S f 2

MFI ¼ Modified foulimg index [s/m6] [s/l2]


Thus, according to Eq. (5.146), there is a relationship between the Modified
Fouling Index and the characteristics of the filter cake, and this index is also
proportional to the particle concentration in the feed to the membrane filter. Because
in practice it is scarcely possible to determine the porosity of the cake deposited on
the filter membrane and the diameter of the particles it is made up of, for the product
of αc and the particle concentration cp,F, an empirical parameter is defined, this being
the fouling index I (Eq. 5.147) [82].

I ¼ αc  cp,F : ð5:147Þ

I ¼ fouling index [m2]


After the MFI has been determined from the gradient of the filtration curve, I can
be calculated from the other test parameters—viscosity μF(t), (S), test pressure Δp,
and filter area Sf—with Eq. (5.147a).

MFI  2  Δp  S f 2
I¼ : ð5:147aÞ
μFðtÞ,ðSÞ

Once the fouling index I is known, the Modified Fouling Index MFI can be
calculated from the other known test parameters with Eq. (5.147b) together with the
flow resistance rc of the membrane coating according to Eq. (5.147c).
510 5 Reverse Osmosis Membrane System: Core Process of SWRO

μF ðtÞ,ðSÞ  I
MFI ¼ : ð5:147bÞ
2  Δp  S f 2
Vf
rc ¼  I: ð5:147cÞ
Sf

MFI0.45 Testing Method16


The Modified Fouling Index MFI0.45 is determined with the same basic testing
equipment as described for measuring the Silt Density Index. In this test, too, the
water under investigation is passed through a flat sheet membrane filter with a mean
pore size of 0.45 μm and a diameter of 47 mm at a constant pressure of 2 bar. The
same flat sheet membrane filter regarding material type, thickness, and flow resis-
tance is used as for the SDI test. However, when measuring the MFI, the
accumulating filtrate volume is noted at uniform brief time intervals. ASTM Stan-
dard D8002, that describes how this test is to be conducted, suggests this time
interval between filtrate volume measurements be not more than every 30 s.
Depending on the fouling potential of the sample under investigation, the filtration
time τf has to be adjusted and the duration of the MFI0.45 test matched to this. The
filtration curve shown in Fig. 5.47 is plotted from the test measurements.
The value of MFI0.45 is calculated from the gradient of the curve in the cake
filtration phase using Eq. (5.148).
τ
Δ V ff
MFI0:45 ¼ ¼ tan α: ð5:148Þ
ΔV f

MFI0.45 ¼ Modified Fouling Index with a 0.45 μm standard filter [s/m6] [s/l2]
τf ¼ filtration time [s]
Vf ¼ filtration volume at filtration time τf
The ASTM standard also includes an algorithm for normalizing the value for
MFItest taken from the filtration curve under the test conditions to the MFISt value at
the standard conditions for viscosity μFðtÞSt ,ðSSt Þ and the test pressure ΔpSt. If a filter
diameter is used that deviates from the 47 mm standard, this normalization equation
(Eq. 5.148a) also permits conversion of the MFI measured under these conditions to
the value that would have been obtained with the standard filter.
  
μFðtÞSt ,ðSSt Þ Δp Sf 2
MFIst ¼ MFItest    : ð5:148aÞ
μFðtÞ,ðSÞ ΔpSt S f ðSt Þ

MFItest ¼ modified fouling index at test conditions [s/l2]


MFISt ¼ modified fouling index at standard conditions [s/l2]
μF(t), (S) ¼ dynamic viscosity of filter feed at test conditions [Pas] [kg/sm]
[Ns/m2]

16
ASTM D8002—Standard Test Method for Modified Fouling Index (MFI-0.45) of Water.
5.3 Fouling and Scaling in RO Systems 511

μFðtÞSt ,ðSSt Þ ¼ dynamic viscosity of filter feed at standard temperature t and stan-
dard salinity S [Pas][kg/sm] [Ns/m2]
Δp ¼ pressure differential at filter at test conditions [Pa] [kg/ms2] [N/m2]
ΔpSt ¼ pressure differential at filter at standard pressure [Pa] [kg/ms2] [N/m2]
Sf ¼ filter area of filter at test conditions [m2]
Sf(St) ¼ filter area of standard filter [m2]

Advantages and Drawbacks of the MFI0.45 Test


The values of the Modified Fouling Index MFI0.45 are mostly directly proportional to
the concentration of the fouling particles in the measured sample. With the help of
the equation for normalization Eq. (5.148a), the values measured for MFI under the
test conditions can be converted to standard values. These two aspects are significant
advantages over the standard SDI approach (SDI drawbacks 1 and 4). Investigations
have shown that the MFI0.45 test may also be employed above the turbidity limit of
<1 NTU as laid down for the SDI test method as well as for surface-extracted water
that has been pretreated by ultrafiltration (SDI drawbacks 5 and 7).
If the MFI0.45 test is done manually, the work required for recording and
evaluating the measured data is greater than is the case for the SDI test.
Regarding testing equipment and membrane filters, the methodology of the
MFI0.45 test as specified in ASTM Standard D800 is the same as the SDI standard
test and accordingly exhibits the same drawbacks regarding its setup and the filter
material that has to be used. Thus, even with the MFI0.45 test, not all fouling
components will be deposited on the membrane (SDI drawback 2). Moreover, the
test results depend greatly on the characteristics of the membrane filter (drawback 4)
and could also be influenced by the sample’s pH (drawback 6).

Equivalence of MFI0.45 and SDI15


The results of these two test methods depend primarily on:

• the flow resistance of the filter material rf


• the filtration pressure Δp
• the area of the filter Sf
• the viscosity of the sample solution μF(t),(S) under consideration of its dependency
on temperature and TDS (in accordance with Sect. 3.2.2.2, Eqs. 3.41, 3.42,
Fig. 3.16, and Table 3.24).

This means that the values from these two test methods may only be considered to
be equivalent if all these parameters are the same when they are carried out. For the
SDI test, additionally the filtration time τf and filtration volume V1 (¼ V2 ¼ Vf) must
be specified.
With knowledge of the above test conditions and parameters, in [96, 97] an
algorithm is developed on the basis of the cake filtration model and corresponding
to Eq. (5.149) with which a theoretical value for SDI15 can be calculated from the
Modified Fouling Index MFI0.45.
512 5 Reverse Osmosis Membrane System: Core Process of SWRO

20
SDI15 ¼
3
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V 22  MFI  ΔP  ∙S f þ V 2  μ2F ðtÞ,ðSÞ  r2f þ 4  τ f ð15Þ  MFI  Δp2  S2f  μ f ðtÞ,ðSÞ  r f  V 1  MFI  V 21  Δp  S f
  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
V 2  V 2  MFI  Δp  S f þ μ2F ðtÞ,ðSÞ  r2f þ 4  τ f ð15Þ  MFI  Δp2  S2f
ð5:149Þ

μF(t),(S) ¼ dynamic viscosity of filter feed at temperature t and salinity S [Pas]


Vf ¼ V1 ¼ V2 ¼ filtration volume for τ1 and τ2 [m3]
MFI ¼ modified fouling index [s/m6]
rf ¼ resistance of filter [m1]
Δp ¼ pressure differential at filter at test conditions [Pa]
τf(15) ¼ elapsed filtration time from start of SDI test up to start of collection of
Volume 2 ¼ 15 min ¼ 900 s [s]
The filtration times τ1 and τ2 in the SDI test are measured for the two equal
filtration volumes V1 and V2. Thus, if in Eq. (5.149) the filtrate volume V1 is set equal
to the volume V2 (¼ Vf), the somewhat simpler Eq. (5.149a) is obtained.

SDI15 ¼ 6:66667
0 1
B μF ðtÞ,ðSÞ r f  MFI  V f  Δp  S f C
 @1  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiA:
V f  MFI  Δp  S f þ μFðtÞ,ðSÞ  r f þ 4  τ f ð15Þ  MFI  Δp  S f
2 2 2 2

ð5:149aÞ

Based on Eq. (5.149a), for an MFI0.45 value of 1.0, Fig. 5.48a shows how the
quotient SDI15/MFI0.45 or the value of the equivalence of the two fouling parameters
varies with the salinity of the water sample under investigation and the test tempera-
ture. These dependencies apply for test filters with a mean flow resistance rf of
1.29  1010 m1. This filter resistance corresponds approximately to a mean value of
the bandwidth of the values for a “pure water flow” of 25 to 50 s/500 ml as specified
in the ASTM standards for the two testing methods. For membrane filters that
deviate by a significant amount either above or below this range, as can be seen
from Eq. (5.149a), the relationship between SDI15 and MFI0.45 is different.
As the graph of Fig. 5.48a shows, with rising temperature the equivalence value
of these two parameters increases, and thus also, for a constant MFI0.45, the value of
SDI15, whereas as salinity increases, and associated with this a rise of viscosity, the
equivalence value and SDI15 decrease. But from Fig. 5.48a, it can be seen that the
influence of temperature on the SDI15/MFI0.45 equivalence value is greater than is
the case for the salinity.
5.3 Fouling and Scaling in RO Systems 513

SDI15 / MFI0.45
4.5

Salinity [g/kg]
20
4.0
50
80
MFI = 1.0*106 [s/m6] = 1.0 [s/l2]
3.5 rf = 1.29*1010 [m-1]

3.0

2.5

2.0

1.5
10 15 20 25 30 35 40 45 50 55
Temperature [°C]

SDI15 / MFI0.45
7
MFI0.45 [s/l2]
0.5
6

Salinity S = 30 [g/kg]
rf = 1.29*1010 [m-1]
5

4 1.0

2.0
2

1
10 15 20 25 30 35 40 45 50 55
Temperature [°C]

Fig. 5.48 (a) SDI/MFI-ratio-dependence on salinity and temperature. (b) SDI/MFI ratio-
dependence on MFI and temperature
514 5 Reverse Osmosis Membrane System: Core Process of SWRO

In the ASTM D8002 standard and also in the membrane manufacturers’ informa-
tion material,17 quoted as an equivalent for the SDI15 target of 3%/min in the feed
to the RO membrane elements is a value for MFI0.45 of 1.0 s/l2. However, the test
conditions needed to attain this relationship are not defined.
Consistent with the dependency of the equivalence value of the two fouling
parameters on temperature and salinity as shown in Fig. 5.48a, for seawater this
relationship of SDI15 to MFI0.45 is roughly attained for a test temperature of 25  C
and a salinity of 20 g/kg, but full agreement is not reached until the temperature is
30  C and the salinity 50 g/kg.
According to Fig. 5.48b, though, the equivalence value of the two fouling
parameters is also fixed by the value of MFI0.45 itself. This figure shows for a salinity
SDI15
of 30 g/kg and MFI0.45 values from 0.5 to 2.0, how MFI 0:45
varies with temperature.
From this, it is apparent for seawater an equivalence value of 3.0 is only attainable for
an MFI0.45 of 1.0 and even then only within a temperature window of 25  C to 30  C.
For values of MFI0.45 and test temperatures that are either side of this, there results a
bandwidth of SDI15/MFI0.45 equivalence values of between 1.5 and over 6.0.
Also for measurements taken at the pretreatment systems of existing SWRO plants,
for which SDI15 and MFI0.45 are found to run parallel over an extended range of the
fouling potential, a similar bandwidth is revealed for the equivalence values of the two
fouling parameters [99, 100]. These data from measurements taken at actual plants,
though, are not enough to obtain correlations between these two parameters derived
from theoretical considerations that would fully confirm the above algorithm.
The ratio of the target values of an SDI15  3%/min to an MFI0.45  1.0 s/l2 as
stipulated in ASTM Standard D8002 and by membrane manufacturers for charging
of RO membrane elements should therefore, as shown above, not be taken as a
generally applicable equivalence value of 3.0 for the two fouling parameters and
used without restrictions for the mutual conversion from one to the other.

Modified Fouling Index-Ultrafiltration, MFIUF


Colloids that form fouling cakes as well as high-molecular-weight and macromolecular
organic substances and organic compounds that promote fouling are not fully detected
with the standard ASTM test methods for SDI15 and MFI0.45. When determining the
fouling potential of a raw water, in order to also take into account the substantial share
of membrane fouling caused by these constituents, the MFI0.45 test method is so
modified that ultrafiltration membranes find application as filter medium [101].
Regarding their retention of particles and high-molecular-weight compounds, UF
membranes are characterized less by their pore size than by their molecular weight
cut-off (MWCO). The extent to which, alongside particulate substances, also high
molecular organic constituents of a sample under investigation are included in the
measurement depends on the MWCO of the selected UF membrane. The lower the
MWCO, the more organic substances with a lesser molecular weight will be
captured by the test membranes and thus be detected as fouling potential.

17
Filmtec Reverse Osmosis Membranes—Technical Manual—2.5.1 Assessment of the Colloidal
Fouling Potential.
5.3 Fouling and Scaling in RO Systems 515

In [101], ultrafiltration membranes with an MWCO range from 1 to 100 kDalton


(kDA) and with differing membrane materials, i.e. polysulphone (PS) and polyacry-
lonitrile (PAN), are investigated regarding their suitability for the MFIUF test. A
13 kDA PAN membrane was proposed for measuring the MFIUF. For further
applications of the MFIUF testing method, like for determining the fouling potential
of seawater prior to pretreatment and after different types of pretreatment processes
in SWRO plants, other UF membranes in a MWCO range of 10–100 kDA and other
membrane materials of different manufacturers (polyether sulfones PES, regenerated
cellulose RC) were used in addition to this reference membrane [102, 103]. It then
becomes apparent that it is not only the MWCO that determines the filtration
behaviour and the capture properties of the UF membranes, but additionally the
membrane material’s physical and chemical structure significantly impacts mem-
brane characteristics and thus also the value of the MFIUF. MFIUF test results should
thus be quoted as a minimum together with the MWCO and material of the
membrane used.
For conducting the MFIUF test under constant pressure, in accordance with the
ASTM standard for MIF0.45, the algorithms for deriving the MFI and the fouling
factor I apply as set forth under Sect. 5.3.3.1.2 (Eqs. 5.145–5.147b), Eq. (5.148) for
determining the MFIUF from the gradient of the filtration curve, and also
Eq. (5.148a) for normalizing the MFIUF from the test conditions.
The MFI test procedure may also be conducted when loading the filter membrane
under constant flux Jf. In order to keep the filtrate flow constant when doing so, the
filtration pressure Δpf is adjusted accordingly [104].
With this type of MFI test, a filtration curve is plotted from the measured data of
the filtration pressure Δpf as it increases over the filtration time τf, while keeping the
flux Jf constant that shows how the filtration pressure Δpf depends on the filtration
time τf. Similar to the curve plotted under constant pressure (see Fig. 5.47), this curve
too shows the phases of blocking filtration, cake filtration, and cake filtration as the
cake becomes compressed. The filtration pressure Δpτ needed to maintain a defined
flux Jf with increasing coating of the filter membrane over time τf is derived from the
series resistance model for filtration in accordance with Eqs. (5.150), (5.150a), and
(5.150b).

Δp f Δp
Jf ¼  ¼  f : ð5:150Þ
μFðtÞ,ðSÞ  r f þ r c μ
Vf
F ðt Þ,ðSÞ  r f þ S f  I

Vf
¼ J f  τf: ð5:150aÞ
Sf

Δpτ ¼ J f  μFðtÞ,ðSÞ  r f þ μFðtÞ,ðSÞ  J f 2  τ f  I: ð5:150bÞ

From the gradient of the Δpf/τf curve in the cake filtration phase, the fouling index
I is calculated in accordance with Eq. (5.150c) or, as shown in Eq. (5.150d), from the
test conditions.
516 5 Reverse Osmosis Membrane System: Core Process of SWRO

ΔΔpτ
Δτ f tan α
I¼ ¼ : ð5:150cÞ
μFðtÞ,ðSÞ  J f 2
μF ðtÞ,ðSÞ  J f 2
Δpτ
μF ðtÞ,ðSÞ Jf rf
I¼ : ð5:150dÞ
J f2  τf

From the value of the fouling Index I, there results the MFItest under test
conditions from Eq. (5.147b) and the normalized value MFISt under standard
conditions with Eq. (5.148a).
Because membrane desalination plants are operated primarily under constant
flux, this type of MFI test conforms better to their operating conditions than is the
case with the standard test under constant pressure. Thus, during the MFI test, it is
possible to match the membrane flux Jf to the values of the flux JW that are usual for
RO plant design [103, 105].
The aim when modifying the MFI testing method in this way is to determine the
parameters for fouling of membranes by particulates under the prevailing test
conditions as well as to predict these. Thus, assuming that the feed pressure Δpτ
needed for a separation membrane is fixed only by the flow resistance rf of the
membrane itself, the build-up of particulate membrane fouling, and the membrane
flux Jf, the increase of this pressure with time as fouling proceeds may be determined
with Eq. (5.151). In order to model “crossflow” operation as is the case for RO
systems, the flow resistance of the cake is multiplied by a fractional deposition
coefficient Ω as; with this mode of operation and depending on flow conditions, only
a portion of the particulate substances are deposited on the membrane (see Sect.
5.3.2.1).

Δpτ ¼ μF ðtÞ,ðSÞ  J f  r f þ μFðtÞ,ðSÞ  Ω  J f 2  τ f  I: ð5:151Þ

Δpτ ¼ feed pressure at filtration or deposition time τ [Pa] [kg/ms2] [N/m2]


Ω ¼ fractional deposition coefficient []
However, this simplifying model cannot be easily applied for RO separation
membranes with high salt rejection and in particular for high seawater salinities. As
described in detail under Sect. 5.3.2.2.2, the fouling deposits decrease the degree of
reverse diffusion of the ionogenic substances from the membrane surface back into
the bulk solution with the consequence that there the elevation of concentration,
i.e. the concentration polarization factor, increases and with it the osmotic pressure at
the membrane surface. For RO seawater desalination, this cake-enhanced osmotic
pressure (CEOP) effect is a major factor leading to the need to raise the feed pressure
to the membrane system as a result of fouling. In comparison, the part played by the
physical pressure increase resulting from the flow resistance presented by the fouling
deposits on the membranes is more on the low side. This impact of the osmotic
effect, CEOP, on the change of pressure with time due to fouling together with the
differing characteristics of the deposits on the membranes of an RO system and those
of the filter cake as revealed by the MFIUF test are to be taken into account by
5.3 Fouling and Scaling in RO Systems 517

augmenting Eq. (5.151) by a cake ratio factor (Eq. 5.151a). Its numerical value is
greater than unity [102, 103].
 
Δpτ ¼ μFðtÞ,ðSÞ  J f  r f þ Ψ  Ω  J f 2  τ f  I : ð5:151aÞ

Ψ ¼ cake ratio factor []


However, the share of the CEOP effect in the cake ratio factor Ψ can only be
determined to sufficient accuracy if the relevant MFI tests are performed using RO
membranes with respective salt rejection characteristics. With UF or NF membranes,
the resulting values of Ψ are too low, so application of Eq. (5.151a) also yields
values for the increase of the feed pressure Δpτ to an RO membrane due to fouling
that are likewise too low.
The extent of deposition of fouling substance onto the surface of a membrane
element and thus also its fractional deposition coefficient Ω may be determined by
drawing up a balance of the concentrations of the foulants in the feed line and their
concentrations in the concentrate emerging from the element, but under consider-
ation of the increase in concentration during desalination, i.e. of the concentration
factor CF or of the recovery rate Y. Because the MSIUF is directly proportional to the
respective concentration of the fouling substance, its concentrations in the feed and
in the concentrate can be set equal to the measured MSIUF values. The fractional
deposition coefficient Ω is then calculated with Eq. (5.151b) [102].
 
1 MFIB 1
Ω¼ þ  1 : ð5:151bÞ
Y MFIF Y
MFIB ¼ MFI of brine/concentrate [s/m6] [s/l2]
MFIF ¼ MFI of feed [s/m6] [s/l2]
Y ¼ recovery factor of membrane element/module/system []
Using Eq. (5.148a), the measured MFIB value must either be converted to the test
conditions of MFIF or else both measured values have to be normalized to standard
MFI conditions with the same equation.
The ‘Modified Fouling Index for Ultrafiltration’ may be used advantageously for
investigating the efficiency of the pretreatment process with regard in particular to
the elimination of fouling substances in the colloidal range and up to the degree of
reduction of higher and high molecular weight substances. It is thus apparent that, for
instance, the MFIUF may also be used as an indicator for biofouling potential in
SWRO plants. Tests with a 10 kDA UF membrane showed a good correlation
between the measured MFIUF10 values and the concentration of assimilable organic
carbon (AOC) in seawater, i.e. of organic substances with a low molecular weight
that initiate and promote bacterial growth [106].
Depending on the type and MWCO bandwidth of the ultrafiltration membranes
used, though, the results obtained are quite specific for the respective testing situa-
tion and the local conditions under which they have been obtained. Although on the
basis of the differing fouling profiles at the various testing locations and pretreatment
518 5 Reverse Osmosis Membrane System: Core Process of SWRO

systems the results obtained are informative regarding the conditions investigated,
usually their numerical values are comparable only to a limited extent.
For investigations and optimization with regard to particulate and colloidal
fouling at existing plants, the MFIUF offers a number of possibilities that go beyond
the potential offered by the Silt Density Index SDI and the MFI0.45. Thus with the
MFIUF, by determining the fractional deposition coefficient Ω at operational
elements, modules or membrane systems, the situation regarding the coating of
their membranes, i.e. the extent of deposition of fouling substances, can be deter-
mined [102, 103, 107]. Also this method may be applied for pilot trials to find the
value of the critical flux Jcrit of a membrane configuration if the fractional deposition
coefficient Ω at various values for membrane flux JW and recovery rate Y are
determined. Thus, the dependency of Ω on both of these factors allows conclusions
to be drawn with regard to the value of the critical flux Jcrit.
Nevertheless, up to now the MFIUF has been employed in the main for pilot tests
and for research investigations at existing plants. One reason for this is that, due to
the elevated flow resistance of UF membranes which also increases with decreasing
MWCO, the testing time is substantially longer compared to the MFI0.45 and may
amount to several hours for each test.

5.3.3.2 Membrane Performance Normalization and Determination


of Fouling Rate
Coating of membrane surfaces by scaling and fouling results in lessening of their
water transport rate and, due to the CEOP effect, also a reduction of salt rejection,
i.e. an increase of salt passage. Resulting in a similar change of membrane element
performance is their ageing with increasing operating time and the associated
so-called irreversible fouling. The impact of both these factors—reversible and
irreversible fouling—on the performance of RO membrane elements can therefore
be established by how much the product flow decreases as well as by the rise in
component concentration or of total dissolved solids (TDS) in the product water
generated by the element.
When operating an RO separation system at constant membrane flux, to compen-
sate for the reduction in product flow and the increasing pressure loss in the flow
channels of the membrane elements due to fouling, the feed pressure to the mem-
brane system is raised accordingly. For this mode of operation that is usually the case
for RO desalination systems, it is then the amount of increase of the feed pressure
together with the elevation of the concentration of the saline components or the TDS
in the product water that characterizes the degree of membrane fouling.
However, like for irreversible and reversible fouling, also fluctuations in the TDS
and temperature in the feed line as well as changes in the recovery rate of a
membrane element, membrane module, and RO unit influence their product flow
and separation action. So as to distinguish between changes in membrane perfor-
mance due to scaling and fouling from changes in product flow and salt rejection as a
consequence of fluctuating operating conditions as they occur during normal opera-
tion of RO desalination systems, it is therefore necessary to convert the respective
current RO operational and performance data to standard or reference data with the
5.3 Fouling and Scaling in RO Systems 519

aid of normalization algorithms. With these normalized data, it is then possible to


present and analyse on a comparable basis the changes in product flow as well as
operating and differential pressure together with the corresponding changes in salt
passage and product composition.

5.3.3.2.1 Performance Normalization


The actual value for the product flow FP(a) of a membrane system is converted to the
product flow FP(St) under standard, or reference, conditions with Eq. (5.152).
Incorporated into this calculation is the ratio of the net driving pressure under
standard conditions NDP(St) to the actual NDP(a) as well as that of the temperature
correction factors for water transport under standard conditions fTCW(St) and under
actual conditions fTCW(a) (Eq. 5.152a). Equation (5.152) corresponds to the algo-
rithm as defined in ASTM standard D451618 for normalization of the product flow of
RO systems.

ΔpFCðStÞ
pFðStÞ   pPðStÞ  π FCðStÞ þ π PðStÞ f TCWðStÞ
F PðStÞ ¼ 2
  F PðaÞ ð5:152Þ
ΔpFCðaÞ
pFðaÞ  2  pPðaÞ  π FCðaÞ, þ π PðaÞ f TCWðaÞ

NDPðStÞ f TCWðStÞ
F PðStÞ ¼   F PðaÞ : ð5:152aÞ
NDPðaÞ f TCWðaÞ

(St) ¼ standard/reference conditions


(a) ¼ actual conditions
The value of the temperature correction coefficient fTCW for the membranes’
water permeability depends on membrane type and the configuration of the mem-
brane element concerned. For calculating this, most manufacturers of polyamide
spiral-wound modules use an algorithm similar to that of Eq. (5.152b). The mem-
brane factor kFW varies depending on the element’s type and configuration.
h  i
kFW  273:15þt ðStÞ 273:15þtðaÞ
1 1

f TCWðStÞ,ðaÞ ¼ e : ð5:152bÞ

If no specific calculation equation for the temperature coefficient is available for


the membrane element to be normalized, it is proposed in ASTM-D4516 that
Eq. (5.152c) be used both for temperature correction of the water permeability and
for the salt passage (see also Sect. 5.1.5.1.1, Eqs. (5.13d), (5.13e), and Fig. 5.17).

f TCWðStÞ,ðaÞ ¼ 1:03tðaÞ 25 : ð5:152cÞ

For different recovery rates Y(a) and Y(St) under the actual conditions (a) and in the
reference mode (St) the differential pressure ΔpFC(a) of the element, module or
membrane array is normalized by applying Eq. (5.153).

18
ASTM D4516—Standard Practice for Standardizing Reverse Osmosis Performance Data.
520 5 Reverse Osmosis Membrane System: Core Process of SWRO

!b
F WF,C,EðStÞ
ΔpFC,EðStÞ ¼ ΔpFC,EðaÞ 
F WF,C,EðaÞ
2  3b
Y
F WF,EðStÞ  1  E2ðStÞ
¼ ΔpFC,EðaÞ  4 
Y
5 : ð5:153Þ
F WF,EðaÞ  1  E2ðaÞ

b ¼ element-specific exponent ¼ 1.3–1.7


ΔpFC, E ¼ pressure loss of element [bar]
F WF,C,E ¼ average feed to concentrate flow of element [m3/h]
FWF, E ¼ feed flow of membrane element [m3/h]
Likewise, the actual values of the salt passage SP(a) and the component concen-
tration cPi(a) or the salt concentration cP(a) of the product water of an RO unit have to
be normalized. This is done with Eq. (5.154), which is applied by virtually all
membrane manufacturers in their respective normalization software tools.

F P,EðaÞ f TCSðStÞ cFBðStÞ cFðaÞ


SPðStÞ ¼ SPðaÞ     : ð5:154Þ
F P,EðStÞ f TCSðaÞ cFBðaÞ cF ðStÞ

The mean product flow F P,E ðStÞ,ðaÞ of the membrane element of an RO unit is the
quotient of its product flow FP(St), (a) and the number of membrane elements N ES it
contains (Eq. 5.154a).

F PðStÞ,ðaÞ
F P,E ðStÞ,ðaÞ ¼ : ð5:154aÞ
N ES

F P,EðStÞ,ðaÞ = mean product flow of membrane element at standard/reference


(St) or actual (a) conditions [m3/h, element]
From F P,EðStÞ,ðaÞ, using Eq. (5.154b), the membrane system’s average specific flux
JP, S ∅ , (St), (a) is calculated and Eq. (5.154c) is then obtained with this parameter.

J P,S∅,ðStÞ,ðaÞ  SE
F P,E ðStÞ,ðaÞ ¼ : ð5:154bÞ
1, 000
J P,S∅,ðaÞ f TCSðStÞ cFBðStÞ cFðaÞ
SPðStÞ ¼ SPðaÞ     : ð5:154cÞ
J P,S∅,ðStÞ f TCSðaÞ cFBðaÞ cFðStÞ

From this equation and the relationship for the feed/brine concentration, that is
the mean bulk concentration cFB(St), (a) according to Eqs. (5.154d), (5.154e) is then
derived.

cFBðStÞ,ðaÞ ¼ cFðStÞ,ðaÞ  CFðStÞ,ðaÞ : ð5:154dÞ


5.3 Fouling and Scaling in RO Systems 521

J P,S∅,ðaÞ f TCSðStÞ CFðStÞ


SPðStÞ ¼ SPðaÞ    : ð5:154eÞ
J P,S∅,ðStÞ f TCSðaÞ CFðaÞ

From this, it is apparent that for normalization of the actual salt passage SP(a) to
SP(St), the determining relationships are the one between the average specific flux
values JP, S ∅ , E(St)(a) and temperature coefficients for membrane permeability fTCS
(St), (a) and the one between the mean concentration factors CFðStÞ,ðaÞ under standard
and under reference conditions.
In a similar approach, the component concentration or salt concentration in the
product water is normalized (Eq. 5.154f). In this case, though, the relationship
between the feed concentration cF(St) under standard conditions and cF(a) under the
actual conditions is an additional normalization parameter.

J P,S∅,ðaÞ f TCSðStÞ cFðStÞ CFðStÞ


cPðStÞ ¼ cPðaÞ     : ð5:154fÞ
J P,S∅,ðStÞ f TCSðaÞ cF ðaÞ CFðaÞ

The mean concentration factor CFðStÞ,ðaÞ is calculated either as the logarithmic


mean CFðlÞ according to Eq. (5.50) or also under consideration of the intrinsic salt
rejection Ri of the membrane as CFðintÞ using Eq. (5.49). The respective recovery
rates of the membrane system, i.e. both the actual recovery rate Y(a) and the recovery
rate under standard conditions Y(St), enter the normalization calculation via these
equations.
Also the calculation and the value of the temperature correction factor for salt
transport fTCS depend on the membrane type as well as on the design and structure of
a membrane element. For calculating this coefficient, most manufacturers of poly-
amide spiral-wound modules apply an algorithm as shown in Eq. (5.154g). How-
ever, also the calculation approach of Eq. (5.154h) is to be found in some
manufacturers’ normalization routines.19 The value of the membrane factor kFS is
fixed by the type of membrane element (see Sect. 5.1.5.1.2, Eq. 5.16c and Fig. 5.17).
h  i
kFS  273:15þtðStÞ 273:15þt ðaÞ
1 1

f TCSðStÞ,ðaÞ ¼ e : ð5:154gÞ

f TCSðStÞ,ðaÞ ¼ ekFS ðtðaÞ tðStÞ Þ : ð5:154hÞ

Software for recording operating data is provided by the membrane


manufacturers (see Annex 5.A1, Table 5.36). In these data acquisition and normali-
zation tools, certain reference conditions can be specified. The entered actual
operating data of an RO unit are then normalized using the normalization algorithms
as described on the basis of the set reference parameters. If the start-up data of a
membrane desalination plant are taken as reference for normalization, it will then be

19
Toray normalization methods SCADA embedded calculations.
522 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fouling rate

Constant flux operation mode

Reversible
Feed pressure increase
fouling
Irreversible fouling
(aging)
Deposition
time ∆τd

Reversible
fouling

Salt passage increase

Membrane operation time ∆τ,M,Op

Fig. 5.49 Fouling rate curve of RO system in constant flux operation mode

possible throughout the plant’s operation to monitor and analyse its performance
data and how this is influenced in particular by membrane ageing and the associated
irreversible fouling as well as by reversible fouling and scaling (see Fig. 5.49).
If the performance status of an operational RO system is to be compared with the
system’s design conditions that have been set as reference and the mean membrane
lifetime τMOpØ,S(a), which is equal to the AMLT(a) of the system’s membrane
element at the selected operating time τMOp,S(a), differs from the design AMLT(d )
(¼ AMLR(St)) that was specified for dimensioning the plant, due to the differing
mean membrane ages τMOpØ,S(a) and τMOpØ,S(St), and thus also the diverging degrees
of irreversible membrane fouling, their water and salt permeability likewise differ at
both operating conditions. In such case of comparison of the membrane perfor-
mance, also the influence of the mean membrane lifetime must therefore be taken
into account. For the normalization of the product performance F PðStÞ,τMOp of a
reverse osmosis system with different mean lifetimes, AMLT(a)(St) of the membranes
therefore in Eq. (5.152) for the calculation of the standard/reference net driving
pressure NDP(St) instead of the standard operating pressure pF(St) according to
Eq. (5.155) the operating pressure pPðStÞτMOp at the mean membrane lifetime AMLT
(a) at which the performance comparison is done is to be applied.

ΔpFCðStÞ
NDPðStÞ,τMOp ¼ pPðStÞ,τMOp   pPðStÞ  π FCðStÞ þ π PðStÞ : ð5:155Þ
2
pPðStÞτMOp = Operating pressure at standard conditions and at τMOpØ,S(a) ¼ AMLT(a)
[bar]
5.3 Fouling and Scaling in RO Systems 523

The normalized reverse osmosis inlet pressure pFðStÞ,τMOp at AMLT(a) can be


determined using the membrane design program of the membrane manufacturer by
changing the value of the design AMLT(d ) to the value of the operating AMLT(a).
However, pPðStÞτMOp can also be determined according to Eq. (5.156) from the
working pressure at design conditions pF(St), D with the corresponding value of the
permeability coefficient for water transport f FP,ðStÞ,τMOp at AMLT(a).

f FPðStÞ
pFðStÞ,τMOp ¼ pFðStÞ,D  : ð5:156Þ
f FPðaÞ,τMOp

fFP(St) ¼ Product permeability coefficient correction factor (flow factor) of ele-


ment at standard (reference) conditions (St) []
f FPðaÞ,τMOp ¼ Product permeability coefficient correction factor (flow factor) of
element at standard (reference) conditions (St) and AMLT(a) []
The salt passage at the operating conditions with AMLT(a) can be determined in
the same way with the membrane design program of the membrane manufacturer as
described for the product performance. If the permeability correction coefficient is
used for salt transport, the normalization of the salt passage SPðStÞ,τMOp and the
product concentration or product salt content cPðStÞ,τMOp is carried out according to
Eqs. (5.157) and (5.157a).

J P,S∅ðaÞ,τMOp f TCSðStÞ CFðStÞ,τMOp


SPðStÞ,τMOp ¼ SPðaÞ,τMOp   
J P,S∅ðStÞ,τMOp f TCSðaÞ CFðaÞ,τMOp
f FSiðaÞ,τMOp
 : ð5:157Þ
f FSiðStÞ

J P,S∅ðaÞ,τMOp f TCSðStÞ cF ðStÞ CFðStÞ,τMOp


cPðStÞ,τMOp ¼ cPðaÞ,τMOp    
J P,S∅ðStÞ,τMOp f TCSðaÞ cF ðaÞ CFðaÞ,τMOp
f FSiðaÞ,τMOp
 : ð5:157aÞ
f FSiðStÞ

fFSi(St) ¼ Salt permeability coefficient correction factor of element at standard


(reference) conditions (St) []
f FSiðaÞ,τMOp ¼ Salt permeability coefficient correction factor of element at actual
conditions (a) and operation time τMOp(d ) []
The product permeability coefficient correction factor f FP,τMOp and the salt
permeability coefficient correction factor f FSi,τMOp are calculated as described under
Sect. 5.1.5.2.3 according to Eqs. (5.72) and (5.73) and with the values in Table 5.3
and can also be taken from the diagram in Fig. 5.24 of this chapter.
524 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.3.3.2.2 Determination of Fouling Rate


When operating an RO system at constant membrane flux, its feed pressure pF not
only has to compensate for the reduction of water permeability caused by membrane
ageing and the irreversible fouling associated with this. It also has to make up for the
progressive increase in the membrane coatings with increasing deposition time ΔτD
and as a consequence the resulting pressure loss ΔpC,Δτ due to the deposition coating
as well as the osmotic pressure Δπ CEOP,Δτ that rises due to the increase of the
component concentration ci,M,ΔτD at the membrane wall owing to the CEOP effect.
With the increase of salt concentration ci,M,ΔτD at the membrane wall, also the
component concentration ciP,Δτ in the product water rises and thus likewise the
salt passage SPiΔτ relative to the feed/brine concentration ciFC.
For an RO system operated in constant flux mode, Fig. 5.49 shows the normalized
values for the system feed pressure pF(St) and salt passage SP(St) derived from the
measured operating data as functions of the membrane operating time ΔτMOp.
These plotted curves, though, are idealized. The degree of reversible fouling
depends on the deposition time ΔτD, i.e. on the time intervals between membrane
cleaning operations. These in turn are fixed for the RO system by how far the
pressure rises or the product TDS rises up to the respective cleaning action being
initiated. If the feed concentration and nature of the fouling substances fluctuate, thus
varying also the characteristics of the membrane coatings, the membrane cleaning
intervals have to be adapted accordingly. Also the type and extent of biological
fouling have a very significant influence on membrane behaviour and how fouling
progresses. This means that the increases of feed pressure and salt passage due to
reversible fouling are not as regular as depicted by the two curves in Fig. 5.49.
The degree of irreversible fouling is influenced by the plant’s operating
conditions, for example the frequency and intensity of disinfectant dosing with
oxidizing chemicals in the feed line to the plant, pretreatment efficiency, the fre-
quency and effectiveness of membrane cleaning, and the chemicals used for this
purpose. Accordingly, the fluctuations in the system’s feed pressure and product
quality caused by irreversible fouling could deviate from the idealized curves shown
in Fig. 5.49.
Also having a significant influence on the degree and course of irreversible
fouling are the extent and chronological sequence of membrane replacement (see
Sect. 5.2.3.3.1). As the membrane replacement rate is extended, the RO system’s
average membrane age AMLT drops and thus also the degree of irreversible fouling.
Once the system’s steady-state operating point has been attained when the AMLT no
longer changes or, if so, only slightly, essentially it is just reversible fouling that
generates the variations in the plant’s operating pressure and salt passage. Thus, the
curves in Fig. 5.49 correspond predominantly to the conditions up to establishment
of the steady-state operating point.
Coating of the membranes of an element, membrane module, or RO system
changes their performance parameters by:
5.3 Fouling and Scaling in RO Systems 525

• reducing the product flow FP and, for operation at constant flux JW, by a
corresponding increase of the feed pressure pF
• increasing the differential pressure ΔpFC due to deposits in the flow channels of
the membrane units
• raising the component concentrations cPi or the TDS cP in the product water.

Through comparative measurements of these performance changes taken once at


the fouled membrane system and then by determining the same parameters after they
have been chemically cleaned, it is possible to ascertain the extent of fouling after a
certain deposition time ΔτD. If the solution is admitted to the membrane system for
both tests under the same standard or reference conditions, for example with
seawater of the same composition or with a standard sodium chloride solution at
the same temperature and recovery rate, the measured values for feed pressure pF,
differential pressure ΔpFC, and component concentrations cPi or TDS cP in the
product water can be used directly for determining the degree of fouling. If the test
conditions differ, these measurements will have to be normalized to a selected
reference basis for the test conditions as described under Sect. 5.3.3.2.1. If the
fouling test is not done at the same membrane system, but instead the fouled
membranes are compared with new, clean membranes, normalization will also
have to take into account membrane ageing, as described under Sect. 5.3.3.2.1.
Determination of the influence of membrane coating on product quality is done
by ascertaining the CEOP fouling factor F CEOP,ΔτD from the quotient of the measured
and normalized values of product concentration ci,pðStÞ,ΔτD under fouled conditions
and the product concentration ci, p(St), clean of the clean membranes with Eq. (5.158).
The value for F CEOP,ΔτD so obtained corresponds at the same time to the relationship
of the concentration polarization factor βi,ðStÞ,ΔτD under fouled conditions to that of
the clean membrane βi,ðStÞ,clean (in this regard, see also Sect. 5.3.2.2.2, Eqs. (5.137),
(5.140), (5.140b), and (5.140c).

ci,pðStÞ,ΔτD βi,ðStÞ,ΔτD
ffi F CEOP,ΔτD ffi : ð5:158Þ
ci,pðStÞ,clean β,ðStÞ,clean

F CEOP,ΔτD ¼ CEOP induced average fouling factor at deposition time ΔτD []
ci,p,ðStÞ,ΔτD ¼ concentration of component i in product of element or system at
standard (reference) conditions (St) and deposition time ΔτD [mg/l] [kg/m3]
ci,p,ðStÞ,clean ¼ concentration of component i in product of element or system at
standard (reference) conditions (St) [mg/l] [kg/m3]
βi,ðStÞ,ΔτD ¼ concentration polarization factor of component i of fouled membrane
element or system at standard (reference) conditions (St) and deposition time ΔτD
[]
βi,ðStÞ,clean ¼ concentration polarization factor average of component i of clean
membrane element or system at standard (reference) conditions (St) []
526 5 Reverse Osmosis Membrane System: Core Process of SWRO

If the concentration polarization factor βiðStÞ,clean is known from calculating the


“clean membrane system” (see Sect. 5.1.5.2.2, Eq. 5.55), then with Eq. (5.158a) the
membrane’s concentration polarization factor βiðStÞ,ΔτD under the fouling conditions
at deposition time ΔτD can be found using the CEOP fouling factor F CEOP,ΔτD
according to Sect. 5.3.2.2.2, Eq. (5.140a). From this, if the intrinsic salt passage
SPi(St) under standard conditions is known, the product concentration ci,p,ðStÞ,ΔτD at
the corresponding deposition time ΔτD may be calculated (Eqs. 5.158b and 5.158c).
The average concentration factor CFi,ðStÞ is to be calculated as a logarithmic mean
with Eq. (5.50) or, by taking the membranes’ salt rejection Ri into consideration,
with Eq. (5.49).

βi,ðStÞ,ΔτD ¼ βi,ðStÞ,clean  F CEOP,ΔτD : ð5:158aÞ

ci,pðStÞ,ΔτD ¼ ci,F ðStÞ  CFi,ðStÞ  βi,ðStÞ,ΔτD  SPi,ðStÞ ð5:158bÞ

ci,p,ΔτD ¼ ci,FðStÞ  CFi,ðStÞ  βi,ðStÞ,clean  F CEOP,ΔτD  SPiðStÞ : ð5:158cÞ

CFi,ðStÞ ¼ concentration factor average of component i at standard (reference)


conditions (St) []
SPi, (St) ¼ salt passage of component i at standard (reference) conditions (St) []
In order to determine the influence of membrane fouling on the necessary
membrane feed pressure pF,TΔτD , like for the tests for the influence on product
quality by fouling, first the operating values of a membrane system with fouling
are measured and then, following membrane cleaning, the same operating data are
measured. These measured values are then normalized to reference conditions. In
this case, these are the operating parameters feed pressure pF,T,(St)ΔτD and pF,T(St),clean
to the membrane system and their respective differential pressures ΔpFCðStÞ,fouled ,ΔτD
and ΔpFCðStÞ,clean .
The feed pressure to the membrane system is made up of various pressure
components. With the normalized measured data, there results primarily the feed
pressure pF,T,ðStÞ,ΔτD to the fouled membranes from the pressure pF, T(St), clean for the
clean membranes and the increase of pressure to the membrane system Δp f ,T,ðStÞ,ΔτD
needed due to fouling (Eq. 5.159).

pF,T,ðStÞ,ΔτD ¼ pF,T,ðStÞ,clean þ Δp f ,T,ðStÞ,ΔτD : ð5:159Þ

pF,T,ðStÞ,ΔτD ¼ total feed pressure of fouled membrane element or system at


standard (reference) conditions (St) and deposition time ΔτD [bar]
pF, T(St), clean ¼ total feed pressure of clean membrane element or system at
standard (reference) conditions (St) [bar]
Δp f ,T,ðStÞ,ΔτD ¼ pressure increase of fouled membrane element or system for
compensation of membrane elements differential pressure and system product flow
loss by fouling at standard (reference) conditions (St) and deposition time ΔτD [bar]
5.3 Fouling and Scaling in RO Systems 527

Δp f ,T ðStÞ,ΔτD is made up of the differential pressure ΔpFC,ðStÞ,ΔτD in the flow


channels of the fouled system and the additional pressure ΔpFC,M ðStÞ,ΔτD that is
needed to overcome the flow resistance of the membrane coating (Eq. 5.159a).

Δp f ,T,ðStÞ,ΔτD ¼ ΔpFC,ðStÞ,ΔτD þ ΔpFC,M ðStÞ,ΔτD : ð5:159aÞ

ΔpFC,ðStÞ,ΔτD ¼ pressure differential total of feed to concentrate of flow channel of


fouled element or system at standard (reference) conditions (St) and deposition time
ΔτD [bar]
ΔpFC,M ðStÞ,ΔτD ¼ pressure differential total for permeation loss compensation of
fouled element or system at standard (reference) conditions (St) and deposition time
ΔτD [bar]
The value of ΔpFC,ðStÞ,ΔτD results as the sum of the differential pressure ΔpFC,ðStÞclean
in the flow channels of the clean system and the increase in differential pressure
ΔpFC,fouled ,ðStÞ,ΔτD due to fouling (Eq. 5.159b). For this, the total of ΔpFC,ðStÞ,ΔτD must
be less than the maximum differential pressure ΔpFCmax that is permissible for the
tested element, module, or membrane system as quoted by the membrane
manufacturer.

ΔpFC,ðStÞ,ΔτD ¼ ΔpFC,ðStÞclean þ ΔpFC,fouled ,ðStÞ,ΔτD  ΔpFCmax : ð5:159bÞ

ΔpFC,ðStÞclean ¼ pressure differential of feed to concentrate of flow channel of clean


element or system at standard (reference) conditions (St) [bar]
ΔpFC,fouled ,ðStÞ,ΔτD ¼ pressure differential of feed to concentrate of element or
system due to flow channel fouling at standard (reference) conditions (St) and
deposition time ΔτD [bar]
ΔpFCmax ¼ maximum allowable pressure differential of feed to concentrate of
membrane element or system [bar]
The differential pressure ΔpFC,fouled ,ðStÞ,ΔτD generated by fouling in the flow
channels is then calculated from the differential pressure measured at the fouled
system ΔpFC,ðStÞ,ΔτD minus the differential pressureΔpFCclean ðStÞ that was determined
for the clean membranes (Eq. 5.159c).

ΔpFC,ðStÞfouled ,ΔτD ¼ ΔpFC,ðStÞ,ΔτD  ΔpFC,ðStÞclean : ð5:159cÞ

The share of the pressure rise ΔpFC,M ðStÞ,ΔτD for the fouling system attributable to
the reduction in permeability because of membrane fouling and the increase of
osmotic pressure due to the CEOP effect is calculated from the total pressure
increase Δp f ,T,ðStÞ,ΔτD to compensate for fouling of the membrane system minus
the differential pressure ΔpFC,ðStÞ,ΔτD of the system (Eq. 5.159d).
528 5 Reverse Osmosis Membrane System: Core Process of SWRO

ΔpFC,M ðStÞ,ΔτD ¼ Δp f ,T,ðStÞ,ΔτD  ΔpFC,ðStÞ,ΔτD : ð5:159dÞ

The value of ΔpFC,M ðStÞ,ΔτD is made up of the increase of osmotic pressure


Δπ FC,CEOPðStÞ,ΔτD caused by fouling at the system’s membrane walls (CEOP effect)
plus the equivalent pressure differential ΔpFC,CakeðStÞ,ΔτD caused by the foulant
coating (Eq. 5.159e).

ΔpFC,M ðStÞ,ΔτD ¼ Δπ FC,CEOPðStÞ,ΔτD þ ΔpFC,CakeðStÞ,ΔτD : ð5:159eÞ

Δπ FC,CEOPðStÞ,ΔτD ¼ average cake-induced osmotic pressure CEOP of fouled


element or system at standard (reference) conditions (St) and deposition time ΔτD
[bar]
ΔpFC,CakeðStÞ,ΔτD ¼ average pressure differential of fouling layer of fouled element
or system at standard (reference) conditions (St) and deposition time ΔτD [bar]
The pressure rise ΔpFC,CakeðStÞ,ΔτD due to the foulant coating can be calculated
with Eq. (5.159f) that is derived from Eqs. (5.159e) and (5.141b).

ΔpFC,CakeðStÞ,ΔτD ¼ ΔpFC,M ðStÞ,ΔτD  π F  CFEðStÞ  βclean


 
 F CEOP,ΔτD  1 : ð5:159fÞ

The necessary total feed pressure of the fouled system pF,T,ðStÞ,ΔτD at the deposi-
tion time ΔτD consequently results as shown in Eq. (5.159g) or, if ΔpFC,CakeðStÞ,ΔτD
pF,T,ðStÞ,ΔτD and is therefore negligible, according to Eq. (5.159h).

pF,T,ðStÞ,ΔτD ¼ pF,T,ðStÞ,clean þ ΔpFC,ðStÞfouled ,Δτ D þ Δπ FC,CEOPðStÞ,ΔτD


þ ΔpFC,CakeðStÞ,ΔτD : ð5:159gÞ

pF,T,ðStÞ,ΔτD ffi pF,T,ðStÞ,clean þ ΔpFC,ðStÞfouled ,ΔτD þ π F  CFðStÞ  βclean


 
 F CEOP,ΔτD  1 : ð5:159hÞ

The fouling factor F fT,ΔτD which encompasses the total influence of reversible
fouling due to both the membrane coating and the increase of osmotic pressure is
calculated as the quotient of the feed pressure pF,T,ðStÞ,ΔτD to the fouled membranes as
measured by the fouling test and then normalized and the normalized pressure
pF, T(St), clean of the clean membranes (Eq. 5.160).

pF,T,ðStÞ,ΔτD
F fT,ΔτD ¼ ð5:160Þ
pF,T,ðStÞ,clean

F fT,ΔτD ¼ factor of total reversible fouling at deposition time ΔτD []


5.3 Fouling and Scaling in RO Systems 529

5.3.3.3 Allowances for Fouling in Membrane Design


The impact on the RO membranes’ water and salt permeability of irreversible
fouling is already considered when preparing the design of the “clean” membrane
system through the respective correction factors f FP,τMOp and f FSi,τMOp for water and
salt fluxes. Depending on the selected pretreatment process, with the membrane
manufacturers’ design software tools, the respective factors for reduction of product
permeability fΔA and increase of salt passage fΔB for the RO system to be dimen-
sioned can be adjusted (see Table 5.3) or they can be matched to the system’s
specific operating situation.
However, when designing an RO system in practice, it is also necessary to take
into account the changes of feed pressure and product concentration due to reversible
fouling. For this purpose, the conditions specified for design of the “clean”
membranes (see Table 5.10) are modified using the fouling factors F CEOP,ΔτD and
F fT,ΔτD determined in Sect. 5.3.3.2.2. The impact of fouling on both these operating
parameters is greatest at the end of the deposition time ΔτD just before a membrane
cleaning operation (see Fig. 5.49) and these values for increase of feed pressure and
of product TDS are to be factored into the considerations depending on the design of
the RO system.
The elevation of salt concentration in the product from a fouled membrane
Δci,p,ðdÞ,ΔτD corresponds to the difference of the concentration polarization factor
βi,ðdÞ,ΔτD that is raised due to the CEOP effect and its value βi,ðdÞ,clean for the clean
membranes (Eq. 5.161).
 
Δci,p,ðdÞ,ΔτD ¼ ci,FðdÞ  CFi,ðdÞ  βi,ðdÞ,ΔτD  βi,ðdÞ,clean  SPi,ðdÞ : ð5:161Þ

Δci,p,ðdÞ,ΔτD ¼ concentration increase of component i in product of element or


system at design conditions (d ) and deposition time ΔτD [mg/l] [kg/m3]
By applying the CEOP fouling factor F CEOP,ΔτD found for the time instant ΔτD as
set out in Sect. 5.3.3.2.2 and Eq. (5.158), it is then possible to determine the
component concentration Δci,p,ðdÞ,ΔτD or the TDS Δcp,ðdÞ,ΔτD of the product water
under design conditions (suffix d ) and at the deposition time ΔτD as shown in
Eq. (5.161a).
 
Δci,pðdÞ,ΔτD ¼ ci,FðdÞ  CFi,ðdÞ  βi,ðdÞ,clean  F CEOP,ΔτD  1  SPi,ðdÞ : ð5:161aÞ

Correction of the feed pressure to the RO system as necessary to take into account
the fouling conditions is done using the fouling factor F fT,ΔτD (Eq. 5.160). This
fouling factor yields the increase of the feed pressure ΔpF,T ðdÞ,ΔτD caused by
reversible fouling at deposition time ΔτD and under design conditions according to
Eq. (5.162).
 
ΔpF,T ðdÞ,ΔτD ¼ pF,T,ðdÞ,clean  F fT,ΔτD  1 : ð5:162Þ
530 5 Reverse Osmosis Membrane System: Core Process of SWRO

ΔpF,T,ðdÞ,ΔτD ¼ increase of total feed pressure of fouled membrane element or


system at design conditions (d ) and deposition time ΔτD [bar]
The pressure rise ΔpF,T,ðdÞ,ΔτD may also be determined from the contributions of
the pressure components according to Eqs. (5.163) and (5.163a) (see also
Eqs. 5.159c and 5.159e).

ΔpF,T,ðdÞ,ΔτD ¼ ΔpFC,ðdÞfouled ,ΔτD þ Δπ FC,CEOPðdÞ,ΔτD þ ΔpFC,CakeðdÞ,ΔτD : ð5:163Þ


 
ΔpF,T,ðdÞ,ΔτD ¼ ΔpFC,ðdÞfouled ,ΔτD þ π F  CFðdÞ  βclean  F CEOP,ΔτD  1
þ ΔpFC,CakeðdÞ,ΔτD : ð5:163aÞ

If the modified fouling index MFIUF is known or after calculating the fouling
index I from the MFIUF with Eq. (5.147a) as described in [105], the pressure loss due
to the membrane coating ΔpFC,CakeðdÞ,ΔτD may be estimated with Eq. (5.163b) for a
certain deposition time ΔτD (see Sects. 5.3.2.2.1 and 5.3.3.1.2).

ΔpFC,CakeðStÞ,ΔτD ¼ μF ðtÞ,ðSÞ  Ψ  Ω  J f 2  τ f  I  105 : ð5:163bÞ

μF(d )(t), (S) ¼ dynamic viscosity of feed at design conditions (temperature t and
salinity S) [Pas]
Ψ ¼ cake ratio factor []
Ω ¼ fractional deposition coefficient []
JP ¼ product flux [m3/m2s]
ΔτD ¼ deposition time [s]
I ¼ fouling index [m2]
Following conventional pretreatment or UF precleaning, for a test rig filtration
flux of 15 l/m2, h and temperature of 20  C, with a 10 kDA UF membrane seawater
exhibits an MFIUF value of 300 to 500 s/l2 [103, 105]. For this MFIUF range and a
average specific flux of 12 to 17 l/m2, h with which RO seawater desalination
systems are operated (see Table 5.12), for a deposition time ΔτD of 6 months,
i.e. membrane cleaning twice per year, and for the assumption of Ω ¼ 1, i.e. complete
particle separation, values in a range of 0.3 to 1 bar are calculated for the pressure
loss ΔpFC,CakeðStÞ,ΔτD due to the flow resistance of the membrane coatings.
For seawater desalination, the pressure loss of the membrane coatings
ΔpFC,CakeðStÞ,ΔτD is significantly less than for the other pressure-raising components
of the fouling process, in particular that of the osmotic influence of the CEOP effect
Δπ FC,CEOPðdÞ,ΔτD . If the membrane coating component is ignored, Eq. (5.163a) leads
to Eq. (5.163c) for calculation of the total pressure differential ΔpF,T,ðdÞ,ΔτD due to
membrane fouling.
5.3 Fouling and Scaling in RO Systems 531

ΔpF,T,ðdÞ,ΔτD ffi ΔpFC,ðdÞfouled ,ΔτD þ π F  CFðdÞ  βclean


 
 F CEOP,ΔτD  1 : ð5:163cÞ

In line with the recommendations of the membrane manufacturers, the admission


of raw water to an RO unit should be halted and membrane cleaning should
commence when the unit’s normalized salt passage has increased, depending on
the specific manufacturer’s stipulations, to between 5% and 20% over that of the
“clean” membranes. The status of “clean membrane” is to be understood as that of
membranes that have undergone a preceding chemical cleaning process, however the
status of their performance (Permeability and salt rejection) being in accordance to
the RO unit’s average lifetime AMLT of its membranes.
A further limit specified for cessation of operation and triggering of chemical
cleaning is stipulated for the normalized differential pressure ΔpFC(St) of the modules
of an RO unit. This should not increase to more than 10% to 20% of the value for the
“clean” units.
These stipulations limit the degree of fouling and the associated changes of the
membrane system’s performance, while at the same time the condition of the
membranes at the maximum deposition time ΔτD,max is defined. At this point in
time, a corresponding value of the fouling factor induced by CEOP F CEOP,ΔτD is
derived from the maximum permissible increase of the salt passage by applying
Eq. (5.140a). For a maximum increase of the normalized salt passage and of the
product TDS of the fouled membrane over the “clean” membrane within a range of
5% to 20%, this then corresponds, according to Eq. (5.140a), to values for F CEOP,ΔτD
of 1.05 to 1.2.
With these two stipulations for the permissible increases of the differential
pressure ΔpFC(St) and of the salt passage Sp(St), with Eq. (5.163d) the rise of the
feed pressure due to membrane fouling at the end of a deposition period ΔpF,ðdÞ,ΔτD
may be calculated.

ΔpF,ðStÞ,Δτ D ffi ΔpFC,ðStÞ,Δτ D  ΔpFC,ðStÞclean þ π F CF ðStÞ βclean


 
F CEOP,Δτ D  1 ð5:163dÞ

For the desalination of seawater with a TDS of 35 g/l at a temperature of 25  C,


the osmotic pressure π F in the feed line to the RO system is approximately 27 bar
(see Sect. 3.2.3.3, Fig. 3.35), while the feed pressure pF, T, (d ), clean of the “clean”
membrane system with desalination membranes with a higher salt rejection is around
53 bar. At a recovery rate of 45%, a “clean” module with eight elements exhibits a
differential pressure of approximately 1 bar (see Fig. 5.36a). For a mean concentra-
tion polarization coefficient βðdÞ,clean of 1.1 and a pressure loss increase of a
membrane module by a factor of 1.1 to 1.2 as well as values for F CEOP,ΔτD of 1.05
to 1.2, with Eq. (5.163d) a fouling-induced total pressure rise ΔpF,T,ðdÞ,ΔτD may be
calculated to within a range of 3 to 9 bar at deposition time ΔτD,max. This
corresponds, when referred to the feed pressure pF, T, (d ), clean of a “clean” membrane
system of 53 bar, to a pressure rise of 6% to 17% that is needed to compensate for
fouling.
532 5 Reverse Osmosis Membrane System: Core Process of SWRO

As the recovery rate Ys(d ) falls, the share of the differential pressure component
ΔpF,C,T,ðdÞ,ΔτD in the total fouling-induced pressure rise ΔpF,T,ðdÞ,ΔτD of the element,
module, or array increases, whereas the osmotic pressure component
Δπ FC,CEOPðdÞ,ΔτD from the CEOP effect decreases. With a rising recovery rate Ys(d )
and thus increasing mean concentration factor CFðdÞ, in contrast the osmotic compo-
nent governs the value of ΔpF,T,ðdÞ,ΔτD to an ever greater degree.
Corresponding to the rise of the normalized salt passage permitted up to the end
of the deposition time ΔτD,max, the salt concentration in the product water increases
by a factor of 1.05 to 1.2, i.e. by 5% to 20%.
When specifying the operating pressures needed for the RO feed and high-
pressure system, the maximum pressure rise pF,T,ðdÞ,ΔτD , max at the end of the deposi-
tion period ΔτD that is needed to compensate for reversible fouling has to be added to
the feed pressure of the clean membranes. The same applies for the component
concentration and salt concentration Δci,p,ðdÞ,ΔτD, max in the product due to fouling
when dimensioning the capacity needed for a second pass of the RO system.
If the split partial design is selected for the seawater RO desalination system (see
Sect. 5.2.3.2.1), it has to be considered that, due to the preferred deposition of fouling
substances in the front elements of an array, it is there that fouling impacts membrane
performance to a much greater extent than at the other membrane elements. The
fouling factors F FT,ΔτD and F CEOP,ΔτD at the feed end exhibit higher values there than
in the downstream elements. Consequently, the salt passage and thus the product TDS
increases more rapidly and to a greater degree in the foremost split partial flow
relative to that in the tail-end partial flow of an array. This means that for the split
partial configuration, unlike for the conventional design, chemical cleaning of the
membrane system should be oriented to the front part stream attaining the maximum
permissible rise of the normalized salt passage of an array, i.e. the maximum
permissible TDS of the bypass part stream of the second pass of the RO system.

5.3.4 Scaling

Defined as scaling is the formation of a coating on the membrane surfaces of RO


elements that arises due to build-up of the concentration of the dissolved components
during the desalination process to such an extent that the solubility limit of sparingly
soluble constituents—referred to as scalants—is exceeded. Scaling therefore occurs
in a membrane module or an RO array, predominantly at the location in the
generated concentrate, the dissolved components are at their maximum concentra-
tion. This is the case particularly at the module’s tail membrane elements of an RO
array. The decisive factor for the formation and extent of scaling is where the
concentration of the ionogenic components which make up the scalant attains their
highest concentration at the membrane surface.
For seawater desalination, compounds that are designated as scalants are:

• the alkaline earth sulphates calcium sulphate CaSO4, barium sulphate BaSO4, and
strontium sulphate SrSO4
5.3 Fouling and Scaling in RO Systems 533

• calcium carbonate CaCO3


• calcium fluoride CaF2
• additionally, magnesium carbonate MgCO3 and magnesium hydroxide Mg(OH)2
if in a two-pass system the pH in the feed line to the second pass is elevated into
the strongly alkaline range.

Silicic acid and phosphate that have to be considered particularly for brackish
water desalination as scalants normally do not present a scaling risk for seawater
desalination due to their low concentrations in seawater and the lower recovery rate
at which SWRO plants are operated.
However, for the high component concentrations of the seawater desalination
concentrates and their complex make-up consisting of a multiplicity of salt
components, scalants are not generally encountered as pure compounds made up
of two components but as composite compounds with several components. In such a
mixed system, it is the shares of the individual scaling components in the mixed
compound which are decisive for the solubility behaviour of the constituents of the
system (see Sect. 3.2.3.2.2, Eq. 3.110 and Fig. 3.32). Added to this are interactions
between the scale-forming components and the remaining saline components of the
solution system, for example due to the formation of ion pairs and other inter-
molecular forces (see Sects. 3.2.3.1 and 3.2.3.2).

5.3.4.1 Scaling: Calculation and Prediction

5.3.4.1.1 Basic Considerations for Determination of Saturation Potential


and Supersaturation
If in a solution made up of several dissolved components Xi ! x, a low solubility
precipitation product Xz is formed as shown by Eq. (5.164), the component
concentrations in equilibrium between the precipitate and the dissolved components
that have combined to generate it may be calculated with Eq. (5.164a).

ni X i þ n j X j þ . . . nx X x ⇄nz X z,solid : ð5:164Þ


n
mnXii,s  mX j j,s  . . . mnXxx,s ¼ K sp,X z : ð5:164aÞ

mX i ¼ mMi,v ¼ molal/molar concentration of component Xi [mol/kgH2O], [mol/l]]


mX i,s ¼ molal/molar saturation concentration of component Xi [mol/kgH2O],
[mol/l]]
ni ¼ stoichiometric factor
K sp,X z ¼ stoichiometric solubility product of scaling components Xz
For a binary precipitation product, i.e. a product that is formed from two solution
components, Eq. (5.164b) then results.

mnX11,s  mnX22,s ¼ K sp,X 3 : ð5:164bÞ


534 5 Reverse Osmosis Membrane System: Core Process of SWRO

Calculation of the concentration relationships in a solubility equilibrium may be


done using either the stoichiometric solubility product K sp (see Sect. 3.2.3.2.2) or on
the basis of thermodynamic modelling with the thermodynamic solubility product
K 0sp (see Sect. 3.2.3.2.1). The stoichiometric solubility product depends on temper-
 
ature, ionic strength, and solution pressure K sp ¼ f ðT, I, pÞ , while the thermody-
namic solubility
  product depends solely on temperature and pressure
K sp ¼ f ðT, pÞ . For thermodynamic modelling, the influence of the solution’s
0

TDS is included in the form of the ionic strength I in the activity coefficient γ X i
γ X i ¼ f ðI Þ . This is calculated either by means of the Debye-Hückel equation or its
modified form, or by using the algorithms of the specific interaction theory, as
explained in detail in Sect. 3.2.3.2.1.
For thermodynamic modelling, both the thermodynamic solubility product K0sp
and the activity coefficient γ X i of the respective solution components have to be
known and calculated by means of γ X i from the molal component concentrations mX i
of their specific activities aX i (see Eqs. 5.164c and 5.164d).

anX11 ,s  anX22 ,s ¼ K 0sp,X 3 : ð5:164cÞ

aX i ¼ m X i  γ X i : ð5:164dÞ

aX i ¼ activity of component Xi [mol/kgH2O]


γ X i ¼ activity coefficient of component Xi []
K0sp ¼ thermodynamic solubility product
But if the stoichiometric solubility product K sp and its dependency on tempera-
ture, ionic strength and, when operating a system at elevated pressure, also its
dependency on pressure are known, it is easier to calculate the solubility equilibrium
using the respective measured component concentrations mX i .
For stoichiometric modelling, the solubility relationships of the solution
components that form the precipitation products in the RO concentrate may be
characterized by the quotients of the ionic concentration product ICP, i.e. the product
of the respective molar or molal concentration of the solution components X1, X2 to
Xi involved in the precipitation reaction, and the solubility product K sp,c,X 1 ,X 2 of the
precipitation product that is generated. This quotient is termed the saturation ratio SR
and is calculated with Eqs. (5.165) and (5.165a).

ICP
SR ¼ : ð5:165Þ
K sp,X 1 ,X 2

ICP ¼ mnX11  mnX22 : ð5:165aÞ

SR ¼ saturation ratio []


ICP ¼ ion concentration product
K sp,X 1 ,X 2 ¼ stoichiometric solubility product of component X1, X2
5.3 Fouling and Scaling in RO Systems 535

For thermodynamic modelling, the SR is calculated as the quotient of the ion


activity product IAP and the thermodynamic solubility product K 0sp,X 1 ,X 2 . In this case,
the IAP results as the product of the activities aX i of the components Xi.

IAP
SR ¼ : ð5:166Þ
K 0sp,X 1 ,X 2

IAP ¼ ion activity product


When modelling scaling in this way, the literature also quotes, as a measure for
the solubility characteristics of low solubility solution components, the logarithm of
the quotient of the IAP and the thermodynamic solubility product K 0sp,X 1 ,X 2 (see
Eqs. 5.166a and 5.166b). This relationship is termed the saturation index SI.

IAP
SI ¼ log ¼ log SR: ð5:166aÞ
K 0sp,cX 1 ,X 2

SI ¼ log IAP  log K 0sp,X 1 ,X 2 ¼ log IAP þ pK 0sp,X 1 ,X 2 : ð5:166bÞ

SI ¼ saturation index []


A further identifier for characterization of the saturation or supersaturation of a
solution containing components that give rise to scaling is the supersaturation ratio
SSR in accordance with Eq. (5.166c). This also takes into account the stoichiometric
factors of the precipitation reaction.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P Ppffiffiffiffiffiffi
ni ICP ni
SSR ¼ ¼ SR: ð5:166cÞ
K sp,X 1 ,X 2

SSR ¼ supersaturation ratio []


Holding true for the parameters of saturation ratio SR and supersaturation ratio
SSR, if their value equals unity is that, although regarding their scaling potential the
solution is saturated, it is in equilibrium, which means that precipitation and thus also
scaling are not to be expected. If the value of SR or SSR exceeds 1, the solution is
supersaturated, solids will be precipitated, and scaling is to be expected. For values
less than 1, some components are undersaturated and the propensity to remain in
solution predominates. With regard to the saturation index SI, there is solubility
equilibrium if its value is zero. For negative values of SI, the components tend to
remain in solution while for positive values they are supersaturated with the ten-
dency to form scale (see Table 5.17).

Table 5.17 Saturation characterizing factors and related precipitation tendency


SR and SSR value SI value Saturation state
1 0 Saturated, in equilibrium
<1 <0 Undersaturated, no precipitation
>1 >0 Supersaturated, precipitation is possible
536 5 Reverse Osmosis Membrane System: Core Process of SWRO

The value of the saturation concentration mX 3,s of the low solubility compound X3
formed from the two components X1 and X2 after adjustment of the solubility
equilibrium can be calculated using Eq. (5.167).
Pqffiffiffiffiffiffiffiffiffiffiffiffi
ni
K sp,X 3
mX 3,s ¼ : ð5:167Þ
nn11  nn22

mX 3,s ¼ molar/molal saturation concentration of low solubility component X3 in


solubility equilibrium [mol/l] [mol/kgH2O]
This calculation value, though, applies only for the case for which the molar and
molal concentrations of X1 and X2 correspond to the stoichiometric ratio of the two
components for the precipitation reaction, but normally this is not the case in a
solution. Then the component that referred to the stoichiometric ratio is in excess
reduces the solubility of the low solubility compound X3.
For binary compounds for which the stoichiometric factors ni of the components
involved are equal (n1 ¼ n2), for instance the alkaline earth sulphates CaSO4, BaSO4,
and SrSO4 as well as calcium carbonate CaCO3, how a difference in concentration
ΔmX 1 ,X 2 between the two components X1 and X2 influences the saturation concentra-
tion mX 3,s of the low solubility compound X3 can be calculated with Eq. (5.167a).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Δm2X 1 ,X 2 þ 4  K sp,X 3  ΔmX 1 ,X 2
mX 3,s ¼ : ð5:167aÞ
2
ΔmX 1 ,X 2 ¼ difference in concentration of X1 and X2—components [mol/l]
[mol/kgH2O]
With Eq. (5.167b), the possible amount ΔmX 3,prec of supersaturation and in this
case of precipitation of the low-soluble compound X3 is calculated. If the solution is
supersaturated, Eq. (5.167b) gives a positive value for the concentration of X3, prec
formed. If there is no supersaturation, the value for ΔmX 3,prec is negative.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
mX 1 þ mX 2  Δm2X 1 ,X 2 þ 4  K sp,X 3
ΔmX 3,prec ¼ : ð5:167bÞ
2
From the value for the concentration ΔmX 3,prec in case of application of molal
concentration units for mX 1 and mX 2 as well as the solubility product K sp,X 3 with
Eq. (5.167c) and if molar concentration is used for these units with Eq. (5.167d)
supersaturation or undersaturation as volume-related concentration, sp,X 3 can be
calculated.
5.3 Fouling and Scaling in RO Systems 537

106  TDSM
sp,X 3 ¼ ΔmX 3,prec  ρs  MWi  : ð5:167cÞ
103
sp,X 3 ¼ ΔmX 3,prec  MWi  103 : ð5:167dÞ

sp,X 3 ¼ amount of precipitating solid [mg/l]


ρs ¼ density of sample [kg/l]
MWi ¼ molecular weight of precipitating compound [g/mol]
TDSM,s ¼ total dissolved solids content of sample, mass-based [mg/kg]

5.3.4.1.2 Saturation Potential Calculation in RO Concentrate


So as to be able to determine the extent to which the solubility limits of specific
scale-forming components are reached or exceeded during desalination, their
concentrations mC,M,X i at the membrane wall at the maximum build-up of concentra-
tion in the system have to be known. By applying Eqs. (5.168) and (5.168a), this is
ascertained from the feed concentration mF,X i to the membrane system with known
concentration factors CFXi for each component and from the concentration polariza-
tion factors βE, t of the elements with the highest build-ups of concentration.
Whether molar or molal units of concentration are used for the calculation
depends on the units in which K*sp is quoted. For thermodynamic modelling, it is
always the molality that is used as the unit of concentration.

1  Y RO  SPX i
mC,M,X i ¼ mF,X i  CFX i  βE,t ¼ mF,X i   βE,t : ð5:168Þ
1  Y RO
1  Y RO  ð1  RX i Þ
mC,M,X i ¼ mF,X i   βE,t : ð5:168aÞ
1  Y RO
mC,M,X i ¼ molal/molar concentration of component Xi in concentrate at membrane
[mol/l] [mol/kgH2O],
mF,X i ¼ molal/molar concentration of component Xi in feed [mol/kgH2O], [mol/l]
CFX i ¼ concentration factor of component Xi []
YRO ¼ RO recovery coefficient []
SPX i = RO salt passage factor []
βE, t ¼ concentration polarization factor at tail membrane elements []
In the tail elements of an RO seawater desalination system, the concentration
polarization factor βE, t, clean for “clean” elements exhibits values that are usually
around 1.05 but may be lower. Nevertheless, when calculating the concentration of
the solution components at the membrane wall, the concentration polarization factor
βE, t should not be neglected. If a coating has already formed on the membrane
surface, the concentration polarization there increases by the CEOP fouling factor
F CEOP,ΔτD due to the CEOP effect (see Sect. 5.3.2.2.2, Eq. 5.140c). For the fouled
membrane, the concentration factor of the clean membrane βE, t, clean is then raised
commensurate with the degree of coating to a significantly higher value βE,t ¼
βCEOP,ΔτD . The concentrations of the scaling components at the membrane surface
and thus also the scaling potential in the RO desalination concentrate increase
538 5 Reverse Osmosis Membrane System: Core Process of SWRO

correspondingly (see Sect. 5.3.2.2.2, Eq. 5.141). Scale that is already present and
incipient scaling promote and speed up further scale formation and thus, under
membrane desalination operation in practice and with membrane conditions that
do not match the “clean” ideal, scaling potential increases accordingly.
Because of the dependency of component solubility or, with stoichiometric
modelling, the stoichiometric solubility product K sp,X z of the scaling products as
well as, for thermodynamic modelling, that of the activity coefficients γ X i of the
solution components on the ionic strength Im,C,M in the concentrate, this must also be
determined at the maximum build-up of concentration. This is done like for deter-
mination of the component concentration mC,M,X i at the membrane wall from the
ionic strength of the feed Im,F, as shown by Eq. (5.168b).

1  Y RO  ð1  RS Þ
I m,C,M ¼ I m,F   βE,t
1  Y RO
1  Y RO  SPS
¼ I m,F   βE,t : ð5:168bÞ
1  Y RO
Im, C, M ¼ ionic strength in concentrate at membrane [mol/kgH2O]
Im, F ¼ ionic strength in feed [mol/kgH2O]
For seawater membranes, the reduction of the ionic strength due to the salt
passage through the membranes is negligible compared to its value in the RO
concentrate, so Eq. (5.168b) is then simplified to Eq. (5.168c).

1
I m,C,M ¼ I m,F   βE,t : ð5:168cÞ
1  Y RO
The ion concentration product ICPC,M in the concentrate at the membrane surface
results from the ion concentration product ICPF in the feed, the concentration factors
CFX i , of the components, and the concentration polarization factor βE,t as shown in
Eqs. (5.169)–(5.169d).

ICPC,M ¼ mnC,M,X
1
1
 mnC,M,X
2
2
: ð5:169Þ
 n  n
ICPC,M ¼ mF,X 1  CFX 1  βE,t 1  mF,X 2  CFX 2  βE,t 2 : ð5:169aÞ
 n  n
ICPC,M ¼ ðmF,X 1 Þn1  ðmF,X 2 Þn2  CFX 1  βE,t 1  CFX 2  βE,t 2 : ð5:169bÞ

ICPF ¼ ðmF,X 1 Þn1  ðmF,X 2 Þn2 : ð5:169cÞ

ICPC,M ¼ ICPF  CFX 1 n1  CFX 2 n2  βE,t n1 þn2 : ð5:169dÞ

ICPC,M ¼ ion concentration product in concentrate at membrane wall


Assuming the same concentration factors CFX i , for the components Xi,
Eq. (5.169d) is simplified to Eq. (5.169e).
5.3 Fouling and Scaling in RO Systems 539

 n þn
ICPC,M = ICPF  CFnX11þn
,X 2  β E,t
2 n1 þn2
= ICPF  CFX 1 ,X 2  βE,t 1 2 : ð5:169eÞ

The saturation ratio SR for the concentrate of an RO desalination system is then


calculated from its ion concentration product ICPC,M at the membranes and the
stoichiometric solubility product K sp,c,X 1 ,X 2 under the conditions of the concentrate
with Eq. (5.170).

ICPC,M
SR ¼ : ð5:170Þ
K sp,c,X 1 ,X 2

The other parameters of the saturation index SI and of the supersaturation ratio
SSR described under Sect. 5.3.4.1.1 together with the calculations outlined in this
section for supersaturation can likewise be applied similar to the saturation ratio SR
on the basis of the algorithms set out there for determination of the scaling potential
in RO concentrates.
The order of calculation for determining the scaling potential of a scaling
component in the concentrate of an RO system on the basis of the scaling ratio is
shown in the flowchart of Table 5.18.
Table 5.18 Sequence of calculation of scaling potentials of scale forming components in RO
concentrate
Equation for
Step Available settings Parameter to be calculated calculation
1 Ionic strength of feed
1.1 Feed analysis mF,X i of all • Ionic strength feed Im,F 3.63
components
2 Ionic strength in concentrate at membrane wall
2.1 Estimate of βE, t or its calculation • Concentration Table 5.10
sequence as per Table 5.10 polarization factor at tail
element βE, t
2.2 Estimate of or RS or SPS, or • Salt rejection or passage Table 5.10
calculation sequences as per RS or SPS
Table 5.10
2.3 Ionic strength feed Im,F, YRO, RS, • Ionic strength of (5.168b), (5.168c)
SPS concentrate Im, C, M
3 Concentration of component Xi in concentrate at membrane wall
3.2 Estimate of RX i or SPX i or their • Salt rejection or passage Table 5.10
calculation sequences as per RX i or SPX i
Table 5.10
3.3 YRO, RX i or SPX i • Concentration factor (5.36), (5.36a)
CFX i
3.4 mF,X i , CFXi or YRO, RX i or SPX i , βE, t • Concentration at (5.168), (5.168a)
membrane wall mC,M,X i
4 Ion product ICPC in concentrate at membrane wall
4.1 mC,M,X i , stoichiometric factors ni • Ion product in (5.169e)
concentrate ICPC,M
5 Saturation ratio in concentrate
5.1 K sp,X 1 ,X 2 , feed temperature tF, Im, C, M • K sp,C,X 1 ,X 2 K sp,X 1 ,X 2 —
Equations in Sect.
5.3.4.1.3
5.2 ICPC,M, K sp,C,X 1 ,X 2 • Saturation ratio SR (5.170)
540 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.3.4.1.3 Stoichiometric Modelling: Stoichiometric Solubility Products


of Scalants: Stiff and Davis Stability Index
From the start of the industrial-scale application of RO membrane technology for
desalination, calculation approaches have been developed by membrane
manufacturers for assessing, reducing, and preventing scaling when using their
products, in particular by DuPont for their Permasep RO elements. The basis for
these calculation procedures was to identify the scaling potential using data for the
solubility limits of low solubility pure binary compounds that could be formed from
the ionogenic constituents of brackish water and seawater.

Such pure scaling compounds in seawater are, as shown in Table 5.19, alkaline
earth sulphates, calcium carbonate, calcium fluoride and, for an SWRO operated in
an alkaline mode in its second pass, also magnesium hydroxide and magnesium
carbonate.
Depending on the composition of the concentrates generated in the desalination
process and the solubility product of the scalants, the compounds with the lowest
solubility determine the possible product recovery of the reverse osmosis process.
The calculation methods are based on stoichiometric modelling of scaling and
corresponding data and graphs for the stoichiometric solubility products of the
named scalants with functions of ionic strength and temperature are provided by
the membrane manufacturers in their engineering and design information material.
For seawater desalination in particular, algorithms for the stoichiometric solubility
product of the various low solubility compounds in seawater as identified in Ocean-
ography research may be used for scaling calculations only to a limited extent as
these mostly have been ascertained up to a maximum salinity of just 40–50 g/kg.
Concentrates from seawater desalination plants, though, may exhibit salinities of
over 70–80 g/kg. The data used by the membrane manufacturers are mainly based on
the determination of the solubility products of the pure compounds of the scalants in
synthetic seawater and sodium chloride solutions, but also in injection waters of oil
fields, where highly saline water is injected into the oil wells to increase the oil yield.
Descriptions of these calculation methods to determine the scaling potential of
alkaline earth sulphates and calcium carbonate as well as provision of the needed
data, graphs, and algorithms to identify the stoichiometric solubility products have
been incorporated into the corresponding ASTM standards, these being:
5.3 Fouling and Scaling in RO Systems 541

Table 5.19 Scalants in seawater and their stoichiometric solubility products


Stoichiometric solubility
products at
temperature ¼ 25  C, ionic
strength ¼ 1 mol/kg H2O, and
Precipitation equation pressure ¼ 1 bar
Scalants in seawater n1  X1 + n2  X2 ⇄ n3  X3 K*sp pK*sp
9
Barium sulphate BaSO4 n1 ¼ n2 ¼ n3 ¼ 1 7.7  10 –9.2  ~8.0–
109 8.1
Calcium sulphate CaSO4 ~1.8  103 ~2.7–
2.8
Strontium sulphate SrSO4 1.4  105–2.4  ~4.6–
105 4.9
Calcium carbonate CaCO3 ~7.1  107 ~6.2
(calcite)
Calcium fluoride CaF2 n1 ¼ 1; n2 ¼ 2; n3 ¼ 1 ~3.3  1010 ~9.5
Magnesium hydroxide Mg ~6.4  1011 ~10.2
(OH)2

• ASTM D4328 and ASTM D4692 for calculating the scaling potential of the
alkaline earth sulphates20

and

• ASTM D3739 and ASTM D4582 for estimating carbonate scaling in brackish
water and seawater together with their concentrates.21

The data and graphs available in the ASTM standards and also in other technical
literature for determining the stoichiometric solubility products of seawater scalants
to then use the values obtained as calculation parameters for estimating scaling
potentials with the algorithms set forth under Sect. 5.3.4.1 are collected and
commented on in the following section.

Alkaline Earth Sulphates: Calcium Sulphate, Barium Sulphate, and Strontium


Sulphate
Calcium Sulphate CaSO4
The graph of Fig. 5.50, for which data from ASTM D4592 are plotted, shows how
the stoichiometric solubility product of CaSO4 K SP,CaSO4 varies with ionic strength Im

20
ASTM D4328 Standard Practice for Calculation of Supersaturation of Barium Sulfate, Strontium
Sulfate, and Calcium Sulfate Dihydrate (Gypsum) in Brackish Waters, Seawater, and Brines;
ASTM D4692 Standard Practice for Calculation and Adjustment of Sulfate Scaling Salts (CaSO4,
SrSO4. and BaSO4) for Reverse Osmosis and Nanofiltration.
21
ASTM D3739 Standard Practice for Calculation and Adjustment of the Langelier Saturation
Index for Reverse Osmosis; ASTM D4582 Standard Practice for Calculation and Adjustment of the
Stiff and Davis Stability Index for Reverse Osmosis.
542 5 Reverse Osmosis Membrane System: Core Process of SWRO

up to a maximum value of 2 for a temperature of 25  C and for data from ASTM


B4328 up to an ionic strength Im of 5 for temperatures from 10  C to 50  C. As the
shape of the curve reveals, within the range of ionic strength of ASTM D4592, the
solubility product K SP,CaSO4 is scarcely affected by temperature. For this range of Im
up to 2, K SP,CaSO4 may be calculated with Eq. (5.171) and as the curve of Fig. 5.50
shows, it is possible to apply the values obtained with this equation up to an ionic
strength Im of 2 not only for a temperature of 25  C but also over the entire
temperature range from 10  C to 50  C.
To be borne in mind is that K SP,CaSO4 in ASTM D4692 is quoted in molar
concentration units, whereas in ASTM D4328 it is as a molal concentration.

K SP,CaSO4 ¼ 4:3224795  105  I 3m  4:2013261  104  I 2m


þ 2:0115484  103  I m þ 1:1186402  104 : ð5:171Þ

Range of validity: Im ¼ 0.1 to 2.0 at 25  C


Barium sulphate BaSO4 and Strontium sulphate SrSO4
The graph in Fig. 5.50 for data from ASTM D4328 shows a comparison of the
solubility product K SP,BaSO4 of barium sulphate BaSO4 in the range of ionic
strength Im of up to 4.5 and a temperature range from 25  C to 50  C and with
data from ASTM D4692 up to an ionic strength Im of 1.5 at a temperature of 25  C. It
is apparent that the curves for a temperature of 25  C differ between these two

K*sp of CaSO4
[x 10-4]
35

30

25 25°C
50°C
ASTM D4692 - molar 35°C
ASTM D4328 - molal
10°C
20

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
Ionic strength [mol/kg H2O]

Fig. 5.50 Stoichiometric solubility product of CaSO4 vs. ionic strength using data from ASTM
D4328 and ASTM D4692
5.3 Fouling and Scaling in RO Systems 543

K*sp of BaSO4
[x 10-9]
50

45 50°C

40

35 35°C
ASTM D4328 - molal

30
25°C

25

20

15

10 25°C

5 ASTM D4692 - molar

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Ionic strength [mol/kg H2O]

Fig. 5.51 Stoichiometric solubility product of BaSO4 vs. ionic strength from ASTM D4328 and
ASTM D4692

guidelines. For BaSO4, too, the units for the two solubility products differ between
the two guidelines as Fig. 5.51 shows.
K SP,BaSO4 can be calculated as a function of ionic strength with Eq. (5.172) with
differing coefficients of the polynomial for ASTM D4692 and for ASTM D4328 at
the corresponding temperatures.

K SP,BaSO4 ¼ a1  I 3m  a2  I 2m þ a3  I m þ a4 : ð5:172Þ

Temperature ( C) a1 a2 a3 a4
ASTM D4692
25 1.28274  1010 9.35734  1010 8.33988  109 1.63442  1010
ASTM D4328
25 1.59266  1010 1.53123  109 9.86718  109 7.03757  1010
35 2.74001  1010 2.41882  109 1.27509  108 8.29581  1010
50 2.92773  1010 3.12541  109 1.80925  108 1.12380  109
Validity: ASTM D4692 from Im ¼ 0.1 to 1.5 at 25  C
ASTM D4328 from Im ¼ 0.1 to 4.5
544 5 Reverse Osmosis Membrane System: Core Process of SWRO

K*sp of SrSO4
[x 10-6]
30

25 °C
25
ASTM D4692-01 - molal
40°C

20
ASTM D4328-08 - molar

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Ionic strength [mol/kg H2O]

Fig. 5.52 Stoichiometric solubility product of SrSO4 vs. ionic strength from ASTM D4328 and
ASTM D4692

The graph of Fig. 5.52 shows how the stoichiometric solubility product K SP,SrSO4

of strontium sulphate SrSO4 varies with ionic strength according to ASTM D4328
over an Im range of up to 2.0 at 25  C and, according to ASTM D4692, up to an Im
value of 3.5 at 40  C. K SP,SrSO4 as a function of ionic strength, pressure, and
temperature can, according to ASTM D 4328, also be calculated using Eq. (5.173)
[108]. However, the range of validity of this algorithm does not commence until a
temperature of 38  C, so that its applicability for lower temperatures is quite limited.
Nevertheless, it does permit calculation of the influence of elevated pressure on the
solubility product of SrSO4. For the pressures at which RO seawater desalination
plants are operated in particular which can reach even up to 80 bar, this pressure
dependency has to be taken into account for scaling calculations.

X
log K SP,SrSO4 ¼ : ð5:173Þ
R
1
X¼ :
T
5.3 Fouling and Scaling in RO Systems 545

pffiffiffiffiffi  2
p
R¼AþBXþC Im þ D  Im þ E  þFX
0:06894757
p pffiffiffiffiffi p
 þ G  Im  :
0:06894757 0:06894757

A +0.266498  103
B 244.828  103
C 0.191065  103
D +53.543  106
E 1.383  1012
F +1.103323  106
G 0.509  109
Validity ranges: Im ¼ 0–3.43, t ¼ 38–149  C, p ¼ 6.9–207 bar
p ¼ pressure [bar]
T ¼ temperature [K]

Also with the algorithm of Eq. (5.174), for both BaSO4 and SrSO4, the pressure
dependency of their respective stoichiometric solubility products K SP,BaSO4 and
K SP,SrSO4 can be calculated from atmospheric pressure up to 300 to 500 bar, starting
from a temperature of 25  C and over an Im range from 0.5 to 4 [mol/l]. In doing so, it
must be borne in mind that in this algorithm the unit used for ionic strength is the
molar concentration referred to NaCl [109, 110].
The graphs in Figs. 5.53 and 5.54 show, for a constant temperature of 25  C and
for an ionic strength I from 0.5 to 4.0 mol/l, the influence of pressure in a range of

K*sp of BaSO4 at 25°C


[x 10-9]
75
Pressure
70 [bar]
65 100
60

55

50 80
45
60
40

35 40
20
30 10
1
25

20

15

10

5
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
NaCl Ionic strength [mol/l]

Fig. 5.53 Stoichiometric solubility product of BaSO4 vs. ionic strength and pressure at 25  C
546 5 Reverse Osmosis Membrane System: Core Process of SWRO

K*sp of SrSO4 at 25°C


[x 10-6]
25

23
Pressure
[bar]
21
100
80
19 60
40
17 20
10
1
15

13

11

5
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
NaCl Ionic strength [mol/l]

Fig. 5.54 Stoichiometric solubility product of SrSO4 vs. ionic strength and pressure at 25  C

1–100 bar on K SP,BaSO4 and K SP,SrSO4 . From the shapes of the curves, it can be seen
that the solubility of barium sulphate in particular increases significantly as the
pressure rises, whereas the influence of pressure on the solubility product of stron-
tium sulphate is much less.
pffiffi pffiffi
ln K SP,SrSO4 ,BaSO4 ¼ A1 þ A2  I þ A3  I þ A4  t þ A5  I p
pffiffi
þ A6  I  t þ A7  I  t þ A8  p  t þ A9
pffiffi
 p2  t þ A10  I  t þ A11  I  ln t þ A12
pffiffi pffiffi
 I  p  ln t þ A13  I  p  ln t þ A14  I
pffiffi
 p2  ln t þ A15  I  p2  ln t þ A16  I  t
pffiffi
 ln t þ A17  I  t  ln t þ A18  I  p  t
pffiffi
 ln t þ A19  I  p  t  ln t þ A20  I  p2  t
 ln t þ A21  I  p2  t  ln t ½mol=l ð5:174Þ
5.3 Fouling and Scaling in RO Systems 547

SrSO4 BaSO4
A1 13.58731 A12 +0.0004058 A1 20.25627 A12 0.001603
A2 +3.336237 A13 0.0001886 A2 0.6153411 A13 0.0004963
A3 0.9129467 A14 +8.145  A3 +2.018038 A14 3.669 
107 106
A4 0.0134939 A15 2.895  A4 +0.0010746 A15 +1.26  106
107
A5 0.0001277 A16 0.0090752 A5 +0.0129667 A16 0.0383439
A6 5.099541 A17 +0.0045206 A6 4.803599 A17 +0.0099981
A7 0.0231996 A18 +4.733  A7 0.0498531 A18 +0.0001036
106
A8 7.032  A19 1.446  A8 0.0002753 A19 3.523 
106 106 105
A9 2.136  A20 9.771  A9 +5.472  A20 1.898 
109 109 107 107
A10 +5.150459 A21 +3.866  A10 +5.026401 A21 +6.625 
109 108
A11 0.0152104 A11 0.5629083
Validity range SrSO4: I ¼ 0.5–4 [mol/l], t ¼ 25–150  C, p ¼ 1 to 300 bar
Validity range BaSO4: I ¼ 0.5–4 [mol/l], t ¼ 25–250  C, p ¼ up to 500 bar
I ¼ ionic strength [mol/l]
p ¼ pressure [bar]
t ¼ temperature [ C]

Calcium Fluoride
The dependency of the stoichiometric solubility product of calcium fluoride CaF2 on
the ionic strength Im in the range of 0.1 to 2.0 at 25  C, as provided by a membrane
manufacturer in his design information material for scaling calculations for brackish
water and seawater desalination [111], is shown by the graph of Fig. 5.55. This curve
for the CaF2 solubility product was then taken over by other membrane
manufacturers for their design documentation [112]. Using Eq. (5.175), solubility
product data may be calculated from the ionic strength in accordance with this curve
at a temperature of 25  C.

K SP,CaF2 ¼ 7:9630318  1011  ln I m þ 3:2906473  1010 ½mol=l ð5:175Þ

Range of validity: Im ¼ 0.1 to 2.0 at 25  C


More recent investigations have revealed, though, that calcium fluoride is more
soluble in seawater than is indicated by the curve of Fig. 5.55. According to [113], in
artificial seawater with an Im of 0.7 at 25  C a solubility product for CaF2 of
1.17  109 was measured against a value of 3.0  1010 as found in these
investigations for binary CaF2/NaCl solutions with the same ionic strength. The
presence of magnesium ions in particular apparently raises the solubility of calcium
fluoride. It may therefore be assumed that the CaF2 solubility product at an Im of 0.7
548 5 Reverse Osmosis Membrane System: Core Process of SWRO

K*sp of CaF2
[x 10-10]
4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0
Ionic strength [mol/kg H2O]

Fig. 5.55 Stoichiometric solubility product of CaF2 vs. ionic strength at 25  C

for the curve in Fig. 5.55 corresponds more to the value in a binary CaF2/NaCl
system and the CaF2 solubility limit in seawater is markedly higher.

Magnesium Hydroxide
The graph in Fig. 5.56 shows how the stoichiometric solubility product K sp,MgðOHÞ
2
of magnesium hydroxide Mg(OH)2 varies with ionic strength in the Im range for
from 0.01 to 1.0 over a temperature span of 15  C to 35  C. Within this ionic strength
range are the salinities in the feed line and in the concentrate of the second pass of an
SWRO where magnesium hydroxide scaling could occur if this RO system is
operated in an alkaline mode.
The graphs were calculated on the basis of a polynomial equation for the
thermodynamic solubility product K 0sp,MgðOHÞ of Mg(OH)2 as a function of tempera-
2
ture in accordance with Eq. (5.176) [114] and the relationship shown in Eq. (5.176a)
between the stoichiometric and the thermodynamic solubility product.

 log K 0sp,MgðOHÞ2 ¼ pK 0sp,MgðOHÞ2


¼ 10:918  0:00304  t þ 0:0000563  t 2 : ð5:176Þ

Range of validity: t ¼ 10–90  C


5.3 Fouling and Scaling in RO Systems 549

K*sp of Mg(OH)2
[ x 10-11]
9

Temperature [°C]
7
35
25
6 15

3
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Ionic strength [mol/kg H2O]

Fig. 5.56 Stoichiometric solubility product of Mg(OH)2 vs. ionic strength at various temperatures

K 0sp,MgðOH Þ
K sp,MgðOH Þ2 ¼ 2
½mol=kg H 2 O ð5:176aÞ
γ Mg2þ  γ 2OH 

The required activity coefficients γ Mg2þ for Mg2+ and γ OH for OH within the
specified ionic strength range may be calculated using the Debye-Hückel algorithms.
In this case, the activity coefficients were calculated using the Davies equation (see
Sect. 3.2.3.2.1, Eq. 3.95).
The algorithms for the influence of pH on the solubility of Mg(OH)2 according to
Eq. (5.176d) and the dependency of the magnesium ion concentration cMg2þ on pH
s,ðtÞ,ðI Þ

(Eq. 5.176e) are derived from the basic equation for the solubility equilibrium of the
compound (Eqs. 5.176b) and (5.176c).

mMg2þ
s
 m2OHs ¼ K sp,MgðOHÞ2 ð5:176bÞ

Kw
mOHs ¼ ð5:176cÞ
mHþs
 
log mMg2þ ¼ log K sp,MgðOHÞ2ðtÞ,ðIÞ þ 2  pK wðtÞ,ðI Þ  pH : ð5:176dÞ
s,ðtÞ,ðI Þ
550 5 Reverse Osmosis Membrane System: Core Process of SWRO

 
log K sp,MgðOHÞ þ2 pK wðtÞ,ðI Þ pH
cMg2þ ¼ 2:432  104  10 2ðtÞ,ðI Þ
: ð5:176eÞ
s,ðt Þ,ðI Þ

K wðtÞ,ðI Þ ¼ stoichiometric ion product of water at temperature t and ionic strength


I or salinity [mol2/kg2]
cMg2þ ¼ solubility of magnesium ions at solubility equilibrium at temperature
s,ðtÞ,ðI Þ

t and ionic strength I or salinity [mg/l]


The curve for the magnesium saturation concentration cMg2þ at a temperature of
s,ðtÞ,ðI Þ

25 C that results from Eq. (5.176e) for the upper pH range of pH 9.5 up to pH 11.0 is
shown in the graph of Fig. 5.57.

Calcium Carbonate Scaling


Like for alkaline earth sulphate scaling, for calcium carbonate scaling, too, a binary
precipitation product is generated from two solution components in accordance with
the following reaction equation.

K spðcÞ
Ca2þ þ CO2
3 , CaCO3 ðs,cÞ

The carbonate concentration is fixed by the proportional concentrations as these


arise for the carbonate/bicarbonate/CO2 equilibrium under the prevailing operating

Magnesium ion
saturation concentration
[mg/l]
600

Water / Concentrate TDS


500 300 mg/l

3,000 mg/l
400
Temperature = 25 °C

300

200

100

0
9.5 9.6 9.7 9.8 9.9 10.0 10.1 10.2 10.3 10.4 10.5 10.6 10.7 10.8 10.9 11.0
pH

Fig. 5.57 Magnesium ion saturation concentration dependence on pH at 25  C


5.3 Fouling and Scaling in RO Systems 551

conditions and their mutual interactions as shown in the following reaction equations
(see Sects. 3.2.4 and 3.2.4.1).

K0
CO2ðgÞ þ H2 O , CO2ðaqÞ
K
CO2ðaqÞ þ H2 O , H2 CO3 ¼ CO2
K1
CO2 þ H2 O , Hþ þ HCO
3

K2
HCO þ
3 , H þ CO3
2

Kw
H2 O , Hþ þ OH

For the stoichiometric approach, the corresponding basic algorithms for


modelling the scaling potential result from the reaction processes shown in accor-
dance with Eqs. (5.177), (5.177a), (5.177b), and (5.177c).

mCa2þ  mCO2
3
¼ K sp,CaCO3 ð5:177Þ

mHþ  mHCO3
¼ K1 ð5:177aÞ
mCO2
mHþ  mCO2
3
¼ K2 ð5:177bÞ
mHCO3

mOH  mHþ ¼ K w ð5:177cÞ

K*1 ¼ first stoichiometric dissociation constant of carbonic acid [mol/kg] [mol/kg


H2O]
K*2 ¼ second stoichiometric dissociation constant of carbonic acid [mol/kg]
[mol/kg H2O]
K w ¼ stoichiometric ion product of water [mol2/kg2] [mol2/kg2H2O]
K sp,CaCO3 ¼ stoichiometric solubility product calcium carbonate [mol2/kg2]
[mol2/kg2H2O]mX i ¼ molar or molal concentration of components [mol/kg]
[mol/kg H2O]
Thermodynamic modelling of scaling is done with the thermodynamic solubility
product of calcium carbonate K 0sp,CaCO3 , the thermodynamic dissociation constants of
carbonic acid K 01 and K 02 , the thermodynamic ion product of water K 0w , and the
respective activity coefficients γ Ca2þ , γ CO2 3
, γ HCO3 , γ HCO3 , and γ OH of the
components concerned (Eq. 5.177d).

mCa2þ  γ Ca2þ  mCO2 γ 2 ¼ K 0sp,CaCO3


3  CO3
ð5:177dÞ
552 5 Reverse Osmosis Membrane System: Core Process of SWRO

When a scaling coating forms, the calcium carbonate may appear in a variety of
crystal structures, like calcite, aragonite, or vaterite (see Sect. 3.2.4.4). These differ
in their solubility and thus also in their stoichiometric and thermodynamic solubility
products. Under the temperature and concentration conditions of RO seawater
desalination, scaling calculations are normally performed based on the calcite
modification.
The propensity for CaCO3 scaling is characterized by the saturation index
SI. This is derived, as shown here by way of example for the stoichiometric
model, from the saturation ratio according to Eq. (5.178).

ICP mCa2þ  mCO2


SR ¼ ¼ 3
ð5:178Þ
K sp,CaCO3 K sp,CaCO3

If, in line with Eq. (5.177b), the carbonate concentration mCO2 3


is substituted by
the concentration of bicarbonate mCO2 3
that is present in the carbonic acid equilib-
rium relationship, Eq. (5.178a) is obtained.

mCa2þ  mHCO3  K 2
SR ¼ : ð5:178aÞ
mHþ  K sp,CaCO3

The saturation index SI is defined as the logarithm of the saturation ratio SR and is
then calculated in accordance with Eq. (5.178b) from the prevailing effective pH pHeff
and the saturation pH pHs as this becomes established if the system comprising
calcium carbonate and the carbonic acid components is in saturation equilibrium,
i.e. calcium carbonate is neither precipitated nor goes into solution.

SI ¼ log SR ¼ pHeff  pHs ð5:178bÞ

pHs ¼ saturation (equilibrium) pH


pHeff ¼ effective (in situ) pH of solution
SI ¼ saturation index

Basic Considerations for Stoichiometric Modelling and Prediction of CaCO3


Scaling Potential in RO Concentrate In a state of saturation equilibrium, the
denominator in Eq. (5.178a) can be set equal to the numerator and the hydrogen
ion concentration mHþ —then corresponds to the concentration mHþs in the saturation
equilibrium relationship (Eq. 5.178c). By taking the logarithms of both sides of the
equation, for stoichiometric modelling the relationship for the saturation pH pHs
according to Eq. (5.178d) and the corresponding algorithm for the saturation index
SI according to Eq. (5.178e) are obtained.
5.3 Fouling and Scaling in RO Systems 553

mCa2þ  mHCO3  K 2
mHþs ¼ ð5:178cÞ
K sp,CaCO3

pHs ¼ pmCa2þ þ pmHCO þ pK 2  pK sp,CaCO : ð5:178dÞ


3 3

 
SI ¼ pHeff  pmCa2þ þ pmHCO þ pK 2  pK sp,CaCO : ð5:178eÞ
3 3

For a seawater solution or the concentrate from an RO desalination plant, the


bicarbonate concentration mHCO3 is calculated from the contribution of the total
carbonate alkalinity mAT,C to its total alkalinity mAT . The total alkalinity mAT is
measured by acidimetric titration for which, alongside the carbonate components,
also other acid-neutralizing constituents are measured, like hydroxide alkalinity,
constituents of the components of the boric acid equilibrium relationship (boron
alkalinity mAB ), and other inorganic and organic compounds with an alkaline action,
i.e. the so-called alkalinity mAMAC of anionic acid-neutralizing components at low
concentrations. The total carbonate alkalinity concentration mAT,C also includes the
hydroxyl ion concentration mOH and the hydrogen ion concentration mHþ . To be
understood as carbonate alkalinity mAC is the total carbonate alkalinity mAT,C without
the H+ and OH shares (the OH and H+ alkalinity mAΔ ) (see Sects. 3.2.4, 3.2.4.1,
and 3.2.4.2).
Normally, the share of alkalinity mAMAC that the minimal inorganic components
contributes to the total alkalinity mAT can be neglected for technical calculations and
so from the total alkalinity mAT there then results the total carbonate alkalinity mAT,C
in accordance with Eq. (5.179). mAT,C is made up of the carbonate alkalinity mAC and
the OH and H+ alkalinity mAΔ as expressed by Eq. (5.179b). From mAT,C , the
carbonate alkalinity mAC is found with Eq. (5.179c) and the bicarbonate concentra-
tion mHCO3 is calculated as shown by Eq. (5.179d).

mAT,C ¼ mAT  mAB  mAMAC : ð5:179Þ

mAT,C ¼ mAC þ mOH  mHþ ¼ mHCO3 þ 2  mCO2


3
þ mOH  mHþ : ð5:179bÞ

mAC ¼ mAT,C  mAΔ ¼ mHCO3 þ 2  mCO2


3
: ð5:179cÞ

mHCO3 ¼ mAC  2  mCO2


3
: ð5:179dÞ

mAT ¼ total alkalinity [mol/kg]


mAB ¼ boron alkalinity [mol/kg]
mAT,C ¼ total carbonate alkalinity with [OH] and [H+] ions [mol/kg]
mAC ¼ carbonate alkalinity [mol/kg]
mAΔ ¼ OH and H+ alkalinity [mol/kg]
mAMAC ¼ alkalinity from minor anionic components [mol/kg]
Taking the basic equations for the bicarbonate/carbonate/carbonic acid equilib-
rium (Eqs. 5.177–5.177c) together with the definition of the total carbonate
554 5 Reverse Osmosis Membrane System: Core Process of SWRO

alkalinity mAT,C in accordance with Eq. (5.179b), from the total carbonate alkalinity
mAT,C and the pH pHeff , the bicarbonate concentration mHCO3 can be calculated with
Eq. (5.180). Additionally, the second stoichiometric dissociation constant of the
carbonic acid K*2 and the stoichiometric ion product of the water K w with the
dependencies of both these on temperature and TDS, or ionic strength, of the
seawater or RO concentrate have to be known.

m2Hþ þ mHþ  mAT,C  K w


mHCO3 ¼
2  K 2 þ mHþ
102pHeff þ mAT,C  10pHeff  K w
¼ : ð5:180Þ
2  K 2 þ 10pHeff

The pmHCO value that has to be known for determining the saturation pH is
3
calculated with Eq. (5.180a).
 
pmHCO ¼ log 2  K 2 þ 10pHeff
3
 
 log 102pHeff þ mAT,C  10pHeff  K w : ð5:180aÞ

This method to determine the bicarbonate concentration under consideration of


the mAΔ alkalinity and the share of the carbonate concentration of the total carbonate
alkalinity mAT,C is necessary for estimating the scaling potential in the seawater
desalination stage of an SWRO, in particular if this is operated in the alkaline mode
at a pH of greater than 8.0 and thus with an elevated content of carbonate in the
concentrate.
The influence of these two parameters of the total carbonate alkalinity becomes
particularly noticeable in the second pass of an SWRO if it is operated at highly
alkaline mode of operation within a pH range of up to 10 for the purpose of greater
boron rejection.
For both modes of operation, due to the increased enrichment of boron in the
concentrate it may become necessary also to take into account the boron alkalinity
mAB when determining the saturation pH pHs and the effective pH pHeff , which then
increases the complexity and computational resources needed for the calculation (see
Sects. 3.2.4 and 3.2.4.3).
If the influence of pH is negligible but the carbonate share in the alkalinity has to
be taken into consideration, the bicarbonate concentration mHCO3 can then be found
from the carbonate alkalinity mAC . For this case, from the above basic equations for
carbonic acid equilibrium and Eqs. (5.179c), (5.181) for determining the bicarbonate
concentration mHCO3 and Eq. (5.181a) for the value of pmHCO may be derived.
3
5.3 Fouling and Scaling in RO Systems 555

10pHeff
mHCO3 ¼ mAC  pHeff : ð5:181Þ
10 þ 2  K2
 
pmHCO ¼ pmA þ pHeff þ log 2  K 2 þ 10pHeff : ð5:181aÞ
3 C

The proportion of carbonate in the lime-carbonic acid equilibrium system of


seawater or concentrate is fixed by its pH as well as its salinity and the temperature
(see Sect. 3.2.4, Figs. 3.39 and 3.40). Accordingly, at high salinity and elevated
temperature, the share of carbonate will exceed 10% of the carbonate alkalinity mAC
already at a pH of 7.3–7.5. Therefore, to avoid inaccuracies when calculating the SI
and thus the scaling potential, above this pH range calculation of the saturation pH
pHs and also the effective pH pHeff of a seawater or RO concentrate should be done as
a minimum on the basis of the carbonate alkalinity mAC .
In this case, the saturation pH may also be calculated with Eq. (5.182).

mCa2þ  mAC
pHs ¼ pK2  log 2 : ð5:182Þ
K sp,CaCO3

Below the above quoted pH range of 7.3 to 7.5, that is in the neutral and down into
the acidic range, carbonate makes up such a small proportion of the carbonate
alkalinity that the carbonate concentration mCO2 3
is negligible. The bicarbonate
concentration mHCO3 may then be set equal to the carbonate alkalinity mAC
(mHCO3 ¼ mAC ).
To determine the potential for calcium carbonate scaling in the concentrate of an
RO desalination system, i.e. its saturation index SIC, with Eqs. (5.183) and (5.183a),
the concentrations at the membrane surface of bicarbonate mC,M,HCO3 , carbonate
mC,M,CO23
, and calcium mC,M,Ca2þ at the maximum build-up of concentration have to
be known.

SIC ¼ pHeff,C  pHs,C : ð5:183Þ


 
SIC ¼ pHeff,C  pmC,M,Ca2þ þ pmC,M,HCO þ pK 2,C  pK sp,CaCO ,C : ð5:183aÞ
3 3

Additionally, for calculating the effective pH pHeff,C , the concentration of carbon


dioxide mC,M,CO2 in the concentrate must be known. For the latter, it is assumed that
the separation membrane does not reject carbon dioxide (RCO2 ¼ 0; SPCO2 ¼ 1Þ ,
i.e. the CO2 concentration in the concentrate has the same value as in the RO feed
and also in the permeate (Eq. 5.183b).

mF,CO2 ¼ mC,M,CO2 ¼ mP,CO2 : ð5:183bÞ

From the RO feed concentration mF,X i of these dissolved components, their


concentrations at the membrane can be determined corresponding to the recovery
556 5 Reverse Osmosis Membrane System: Core Process of SWRO

rate YRO of the RO system, the salt passage SPX i of each of the solution components,
and the concentration polarization factor βE, t at the maximum build-up of concen-
tration, as shown in Eq. (5.183c).

1  Y RO  SPX i
mC,M,X i ¼ mF,X i  CFX i  βE,t ¼ mF,X i   βE,t : ð5:183cÞ
1  Y RO

X i ¼ HCO 2
3 ; CO3 ; Ca

Therefore, to calculate the saturation pH pHs,C of the concentrate using


Eq. (5.183a), additionally the second stoichiometric dissociation constant of the
carbon dioxide K 2,C and the solubility product of calcium carbonate K sp,CaCO3 ,C
have to be known for the concentrate’s conditions of salt content and temperature.
When designing an RO system, normally the effective pH pHeff of the seawater is
known from measurements. However, the pHeff,C in the concentrate has to be
calculated from the components of the carbonic acid equilibrium relationship. This
may be done on the basis of the carbonate alkalinity mC,M,AC and the content of CO2
mC,M,CO2 in the concentrate. In accordance with Eq. (5.183d), the carbonate alkalinity
mC,M,AC is calculated from the values for the bicarbonate concentration mC,M,HCO3
and the concentration of carbonate mC,M,CO2 3
as previously calculated with
Eq. (5.183c). The CO2 content in the concentrate corresponds to the value for the
CO2 in the RO feed (mF,CO2 ¼ mC,M,CO2 Þ.

mC,M,AC ¼ mC,M,HCO3 þ 2  mC,M,CO2


3
: ð5:183dÞ

With these data, the effective pH pHeff,C in the concentrate may be determined
using the algorithms of Eqs. (5.184)–(5.184c). Details of the derivation of these
equations from the basic equations of the carbonic acid equilibrium may be found
under Sects. 3.2.4 and 3.2.4.3.
ffi!
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2
p p
pHeff,C ¼  log  þ q : ð5:184Þ
2 2

K 1,C  mC,M,CO2
p¼ : ð5:184aÞ
mC,M,AC
2  mC,M,CO2
q ¼ K 1,C K 2,C  : ð5:184bÞ
mC,M,AC
5.3 Fouling and Scaling in RO Systems 557

0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1

K 1,C  mC,M,CO2 K 1,C  mC,M,CO2 2 2  mC,M,CO2
pHeff,C ¼ log @ þ þ K 1,C  K 2,C  A:
2  mC,M,AC 2  mC,M,AC mC,M,AC

ð5:184cÞ

But in addition to these named concentration values, for calculating pHeff,C also the
first and second stoichiometric dissociation constants of the carbonic acid K 1,C and
K 2,C together with their dependency on the TDS and temperature of the concen-
trate have to be known. Depending on TDS and temperature, below a pHeff,C range
of pHeff,C ¼ <0.0–7.5, due to the then negligibly low mC,M,CO2 3
concentration, this
may calculated with Eq. (5.184d).

mC,M,HCO3
pHeff,C ¼ pK 1,C þ log : ð5:184dÞ
mC,M,CO2

Then, though, only the value of the first dissociation constant K 1,C has to be
known.
Polynomials for K 1 and K 2 as well as the stoichiometric solubility product of the
calcium carbonate K sp,CaCO3 have been determined in oceanographic investigations
for a validity range of salinity of up to 40–50 g/kg seawater and a maximum
temperature of 50  C (see Sects. 3.2.4, 3.2.4.1, and 3.2.4.4). If these algorithms
are applied for concentration calculations above these ranges of validity, the results
obtained however are no longer corroborated by experimental results.
Even in high salinity RO concentrates, the Stiff-Davis calculation method allows
calculation of the saturation pH and, if the effective pH in the concentrate is known,
thus also a saturation index that characterizes the calcium carbonate scaling potential
of the concentrate, termed the Stiff & Davis stability index.

Stiff & Davis Stability Index, S&DSI


The Stiff and Davis calculation procedure has established itself as the standard
method in RO desalination system technology for assessing the calcium carbonate
scaling potential of seawater and highly saline concentrates in RO desalination
system technology. The respective algorithms are presented in ASTM D458222
together with a description of the calculation procedure.
Like the saturation index SI, the S&DSI is calculated from the difference of the
effective pH pHeff and the saturation pH pHs (Eq. 5.185). Unlike the SI calculation,
though, the saturation pH is determined using a constant that is dependent on the
solution’s ionic strength and temperature, the S&DSI stability index constant KS&DSI
(Eq. 5.185a). This is an empirical constant whose values as a function of ionic
strength and temperature were established by H.A. Stiff and E.D. Davis in tests with

22
ASTM D4582 Standard Practice for Calculation and Adjustment of the Stiff and Davis Stability
Index for Reverse Osmosis.
558 5 Reverse Osmosis Membrane System: Core Process of SWRO

NaCl/CaCl2/Na2CO3 solution within an Im range of up to 3.6 mol/kg H2O and


temperatures from 0  C to 50  C [115].
If the saturation pH pHs is calculated in accordance with ASTM D4582, the total
alkalinity mAT is taken as the basis for the carbonate concentration that is relevant for
scaling. However, for determining the S&DSI in RO osmosis concentrates, it proves
to be better to apply the alkalinity as carbonate alkalinity mC,M,AC . This is calculated
from the concentrations mC,M,HCO3 and mC,M,CO2 3
at the membrane wall in line with
Eq. (5.185c). This concentration of the two components results from the determina-
tion of the concentrate and permeate composition during membrane design using
Eq. (5.183c), where also the differing salt passages of the two components then
additionally being taken into consideration.
The S&DSI in the concentrate of an RO desalination system is then calculated
from the algorithms of Eqs. (5.185b)–(5.185e) together with Eq. (5.185f).

S&DSI ¼ pHeff  pHs : ð5:185Þ

pHs ¼ pmCa2þ þ pmAT þ K S&DSI : ð5:185aÞ


 
S&DSIC ¼ pHeff,C  pmC,M,Ca2þ þ pmC,M,A þ K S&DSI,C ð5:185bÞ
C

mC,M,AC ¼ mC,M,HCO3 þ 2  mC,M,CO2


3
ð5:185cÞ

pmC,M,Ca2þ ¼  log mC,M,Ca2þ ð5:185dÞ

pmC,M,AC ¼  log mC,M,AC ð5:185eÞ

S&DSI ¼ Stiff & Davis Stability Index


KS&DSI ¼ Stiff & Davis Stability Index constant
The value of the constant KS&DSI for a specific ionic strength Im in mol/kg H2O
and a temperature t in  C is calculated using a polynomial equation given in ASTM
D4582-10 and as shown by Eq. (5.185f).

K S&DSI ¼ 3:78342 þ 0:16781  ln I m  0:26411  ð ln I m Þ2  0:1029


 ð ln I m Þ3  0:01124  ð ln I m Þ4  0:01221  t  0:0001316
 t2 : ð5:185fÞ

Ranges of validity: I ¼ 0 – 3.6 mol/kg H2O; t ¼ 0 – 50 CThe KS&DSI values are
plotted as functions of ionic strength and temperature in the graph of Fig. 5.58.
In comparison with the calculation of the saturation index, the Stiff & Davis
stability index constant KS&DSI corresponds to the difference of the pK value of the
second dissociation constant of the carbonic acid K 2 and that of the solubility
product of calcium carbonate K sp,CaCO3 , or rather the logarithm of the quotient of
these two parameters according to Eq. (5.185g).
5.3 Fouling and Scaling in RO Systems 559

KS&DI
3.8
3.7
3.6
3.5
3.4 Temperature
[ °C ]
3.3
3.2 10
15
3.1
20
3.0
25
2.9
30
2.8
35
2.7
40
2.6
2.5 45

2.4 50
2.3
2.2
2.1
2.0
0.0 1.0 2.0 3.0 4.0
Ionic strength [mol/kg H2O]

Fig. 5.58 Stiff & Davis Stability Index—KS&DSI coefficient—Dependence on ionic strength and
temperature

K sp,CaCO3
K S&DSI ffi pK 2  pK sp,CaCO ffi log : ð5:185gÞ
3 K2

When designing an RO desalination system, no measured value for the effective


pH in the concentrate is available. This must be calculated from the parameters of the
carbonic acid equilibrium under the specified design conditions. In accordance with
the calculation method of ASTM D4582 for the Stiff & Davis stability index for an
RO concentrate, the effective pH pHeff,C in the concentrate according to Eq. (5.186) is
determined from the total alkalinity there prevailing cC,AC,T as mg/l CaCO3 and the
content of carbonic acid cC,CO2 in mg/l CO2.

cC,AC,T
pHeff,C ¼ 0:423  ln þ 6:2033: ð5:186Þ
cC,CO2

cC,AC,T ¼ total alkalinity in concentrate [mg/l CaCO3]


cC,CO2 ¼ CO2 content in concentrate [mg/l CO2]
Equation (5.186) is based on the simplifying Eq. (5.184d) and it contains a fixed
value for pK 1,C . This means that changes of the salt concentration of the concentrate
and its temperature with their influence on the determination of pHeff,C are not taken
into account. Further, for simplification, the total alkalinity cC,AC,T in mg/l CaCO3 is
equated with the bicarbonate concentration cC,HCO3 in mg/l CaCO3.
560 5 Reverse Osmosis Membrane System: Core Process of SWRO

The influence of the salt concentration of the concentrate and its temperature may
be accounted for if when calculating the pHeff,C Eq. (5.184c) is used and the values of
the first and second dissociation coefficients of the carbonic acid K 1 and K 2 needed
for this purpose with their dependency on ionic strength and temperature are
established using the polynomial relationship Eq. (5.187). The algorithms for this
equation were derived by Millero, Huang et al. [116] from the results of tests with
NaCl/Na2CO3 solutions within a concentration range of 0–6 mol NaCl/kg H2O and
temperatures from 0  C to 50  C.

Bi
pK i ¼ pK 0i þ Ai þ þ Ci  ln ðT Þ ½mol=kgH 2 O: ð5:187Þ
T
Ai ¼ a0  I 0:5 þ a1  I þ a2  I 1:5 þ a3  I 2

Bi ¼ a4  I 0:5

C i ¼ a5  I 0:5
b1
pK 0i ¼ b0 þ þ b2  ln ðT Þ
T

Parameter Coefficients
pK*1 pK*2
a0 +35.2911 +38.2746
a1 +0.8491 +1.6057
a2 0.32 0.647
a3 +0.055 +0.113
a4 1583.09 1738.16
a5 5.4366 6.0346
pK01 pK02
b0 114.3106 83.2997
b1 +5773.67 +4821.38
b2 +17.779524 +13.5962
Ranges of validity: I ¼ 0 – 6 mol/kgH2O; t ¼ 0 – 50  C

The graphs of the two stoichiometric dissociation coefficients K 1 and K 2 are plotted
as functions of the NaCl molality/ionic strength and of the temperature in Fig. 5.59a,
b. The values of the dissociation constants K 1 and K 2 are calculated from their pK i
values using Eq. (5.187a).

K i ¼ 10pK i : ð5:187aÞ

The sequence for calculating the S&DSI in the concentrate of an RO desalination


system is shown in Table 5.20, stating for each calculation step the equations to be
used to determine the parameters.
5.3 Fouling and Scaling in RO Systems 561

K*1[x 10-6]
1.6

1.5
Temperature

1.4
50

1.3 40

1.2
30

1.1 25

1.0 20

0.9
10
0.8

0.7
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
Ionic strength - NaCl molality [mol/kg H2O]

K*2 [x 10-10]
7.0
Temperature

6.0 50

5.0 40

30
4.0
25
20
3.0
10

2.0

1.0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
Ionic strength - NaCl molality [mol/kg H2O]

Fig. 5.59 (a) First stoichiometric dissociation constant of carbonic acid—dependence on NaCl
molality and temperature. (b) Second stoichiometric dissociation constant of carbonic acid—
dependence on NaCl molality and temperature
562 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.20 Sequence of Stiff & Davis Stability Index (S&DSI) calculation in RO concentrate
Equation for
Step Available settings Parameter to be calculated calculation
1 Ionic strength Im,F of feed
1.1 Feed analysis, mF,X i of all components • Ionic strength feed Im,F (3.63)
2 Ionic strength Im,C in concentrate at membrane wall
2.1 Estimate of βE, t or its calculation • Concentration polarization Table 5.10
sequence as per Table 5.10 factor at tail element βE, t
2.2 Estimate of or RS or SPS or calculation • Salt rejection or passage RS Table 5.10
sequence as per Table 5.10 or SPS
2.3 Ionic strength of feed Im,F, YRO, RS, • Ionic strength of concentrate (5.168b),
SPS Im, C, M (5.168c)
3 Concentration of components Xi = Ca2+, HCO32, CO322 in concentrate at
membrane wall
3.1 Estimate of RX i or SPX i or its • Salt rejection or passage RX i Table 5.10
calculation sequence as per Table 5.10 or SPX i
3.2 YRO, RX i or SPX i • Concentration factor CFX i (5.36),
(5.36a)
3.3 mF,X i , CFX i or YRO, RX i or SPX i , βE, t • Membrane wall 5.168),
concentration mC,M,X i 5.168a)
4 Saturation pHs,C in concentrate
4.1 mC,M,HCO3 mC,M,CO2
3
• mC,M,AC 5.185c)
4.2 mC,M,AC • pmC,M,AC 5.185e)
4.3 mC,M,Ca2þ • pmC,M,Ca2þ 5.185d)
4.4 Im, C, M, tC • S & DSI stability index 5.185f)
constant KS & DSI, C
4.5 pmC,M,AC , pmC,M,Ca2þ , KS & DSI, C • Saturation pH in concentrate 5.185a)
pHs,C
5 Effective pHeff,C in concentrate
5.1 Im, C, M, tC • 1st dissociation constant 5.187),
K 1,C 5.187a)
5.2 Im, C, M, tC • 2nd dissociation constant 5.187),
K 2,C 5.187a)
5.3 mF,CO2 ¼ mC,M,CO2 , mC,M,CO2
3
, K 1,C , • Effective pH in concentrate 5.184c)
pHeff,C
K 2,C
6 Stiff & Davis stability index S&DSI in concentrate
6.1 pHeff,C , pHs,C • S&SDIC in concentrate 5.185)

5.3.4.1.4 Thermodynamic Modelling of CaCO3 Scaling: Langelier Saturation


Index
In order to set up a thermodynamic model of the CaCO3 scaling potential, apart from
the concentrations of the reagents Ca2+, HCO3, and H+, also their activity
coefficients γ Ca2þ , γ HCO3 and γ Hþ as well as the second thermodynamic dissociation
5.3 Fouling and Scaling in RO Systems 563

coefficient of the carbonic acid K 0sp,CaCO3 have to be known (see also Sects. 3.2.3, 3.2.
3.2, and 3.2.3.2.1 for further details and a description of thermodynamic modelling
of solution equilibriums under the application of activity coefficients).
From these parameters, the saturation ratio SRTh can be calculated with
Eq. (5.188). The activity coefficients of the components may be combined into an
activity coefficients factor fY by applying Eqs. (5.188a) and (5.188b).

mCa2þ  γ Ca2þ  mHCO3  γ HCO3  K 02


SRTh ¼ : ð5:188Þ
mH þ  γ Hþ  K 0sp,CaCO3
γ Ca2þ  γ HCO3
¼ f γ: ð5:188aÞ
γ Hþ

mCa2þ  mHCO3  K 02
SRTh ¼  f γ: ð5:188bÞ
mHþ  K 0sp,CaCO3

SRTh ¼ saturation ratio with thermodynamic modelling [–]K02 ¼ thermodynamic


dissociation constant of carbonic acid [mol/kg H2O]
K 0sp,CaCO3 ¼ thermodynamic solubility product calcium carbonate [(mol/kg
H2O)2]
γ Ca2þ ¼ activity coefficient Ca2+
γ HCO3 ¼ activity coefficient HCO3
γ CO2
3
¼ activity coefficient CO32
γ Hþ ¼ activity coefficient H+
fY ¼ activity coefficients factor
As shown in Eq. (5.189a) and like for stoichiometric modelling, the thermody-
namic saturation index SITh is derived from the saturation pH pHs in accordance with
Eq. (5.189) and the effective pH pHeff , but in this case with the saturation pH
determined while taking into account the activity coefficient factor fY.

pHs ¼ pK 02  pK 0sp,CaCO þ pmHCO þ pmCa2þ þ p f γ : ð5:189Þ


3 3

SITh ¼ LSI ¼ pHeff  pHs


 
¼ pHeff  pK 02  pK 0sp,CaCO þ pmHCO þ pmCa2þ þ p f γ : ð5:189aÞ
3 3

SITh ¼ saturation index with thermodynamic modelling [–]


LSI ¼ Langelier saturation indexThe bicarbonate concentration mHCO3 for
determination of the CaCO3 scaling potential can be determined from the total
carbonate alkalinity mAT,C with Eq. (5.189b) and its associated value for pmHCO
3
using Eq. (5.189c).
564 5 Reverse Osmosis Membrane System: Core Process of SWRO

2pH
K 0w
10 eff
γ Hþ þ 10pHeff  mAT,C  γ
OH
mHCO3 ¼ 2γ HCO K 02
ð5:189bÞ
pHeff
10 þ 3
γ CO2
3

!
2  γ HCO3  K 2 0
pmHCO ¼ log 10pHeff þ
3 γ CO2
3

102pHeff K0
 log þ 10pHeff  mAT,C  w ð5:189cÞ
γHþ γ OH

The parameters mHCO3 and pmHCO are calculated from the carbonate alkalinity
3
mAC with Eqs. (5.189d) and (5.189e).

10pH  γ CO2
mHCO3 ¼ mAC  3
: ð5:189dÞ
10pH  γ CO2
3
þ 2  K 02  γ HCO3

10pH  γ CO2 þ 2  K 02  γ HCO3


pmHCO ¼ pmA þ log 3
pH : ð5:189eÞ
3 C 10  γ CO2
3

The effective pH pHeff is calculated from the carbonate alkalinity mAC and the CO2
concentration mCO2 as shown by Eq. (5.190).
0 vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1
u !
u mCO  K 01  γ CO 2
B mCO2  K 1  γ CO2  m  γ
0
t 2 CO CO 2 C
pHeff ¼ log @ þ 2 2
þ K 01  K 02  2
A
2  mAC  γ Hþ  γ HCO3 mAC  γ Hþ  γ HCO3 mAC  γ CO2  γ 2Hþ 3

ð5:190Þ

Further explanations regarding the calculation of effective pH pHeff including the


influence of the ion product of the water (from mAT,C ) and also the boric acid/borate
equilibrium (from mA0T ), for example when operating membrane desalination in the
alkaline range, are to be found in Sects. 3.2.4, 3.2.4.1, 3.2.4.2, and 3.2.4.3.
Calculation of the effective pH for the solubility equilibrium range in which the
influence of the carbonate concentration mCO2 3
is negligible may be done with the
simplified Eq. (5.190a).

mHCO3  γ HCO3
pHeff ¼ pK 01 þ log ð5:190aÞ
mCO2  γ CO2

The values of the thermodynamic parameters K 01 , K 02 , and K 0sp,CaCO3 as functions


of temperature T in Kelvin may be calculated according to [117] using the algorithms
of Eq. (5.191) and the coefficients listed there for the respective parameter.
5.3 Fouling and Scaling in RO Systems 565

C E
pX ¼ A þ B  T þ þ D  log T þ 2 : ð5:191Þ
T T

Coefficient
Parameter X A B C D E
K01 356.3094 0.06091964 21,834.37 126.8339 1,684,915
K02 107.8871 0.03252849 5151.79 38.92561 563,713.9
K0sp,CaCO3 171.9065 0.077993 2839.319 71.595 –

K 01
Validity range: and K 02
T ¼ 273–373 K
K 0sp,CaCO3,C T ¼ 273–363 K

The pK value of the ion product of the water pK 0w and its temperature dependency is
calculated with Eq. (5.191a) [118].

4471
pK 0w ¼ þ 0:01706  T  6:0875: ð5:191aÞ
T
Validity range: T ¼ 273–333 K
The activity coefficients of the components participating in the solubility equilib-
rium together with the influence of salt concentration and ionic strength on the value
of these coefficients are determined on the basis of the Debye-Hückel theory
(DH theory) and the algorithms derived from this. For this, up to which TDS the
values of the saturation index SITh obtained from this modelling correspond to the
actual scaling potential of the water depend on the type of algorithm based on the DH
theory used for calculating the activity coefficient (see Sects. 3.2.3 and 3.2.3.2.1,
Table 3.10).

Langelier Saturation Index LSI


The saturation index SITh is named the Langelier Saturation Index LSI after W. F.
Langelier. He was the first to rate by means of a saturation index the tendency of
calcium carbonate to either precipitate or to redissolve in fresh water. He derived this
index from thermodynamic modelling of the calcium carbonate/carbonic acid equi-
librium relationship together with the DH theory for calculation of the activity
coefficient [119]. Later on, his algorithm, that initially was only applicable for
low-salinity water, was adapted for application to water of higher salinity by
modifying the DH algorithm [120].

Standards for SI and LSI Calculation For this calculation model for assessing the
corrosivity and tendency to form scale of drinking water, a number of standards and
guidelines have been drawn up in each of which the theoretical basis is set forth and
the calculation procedures explained.
566 5 Reverse Osmosis Membrane System: Core Process of SWRO

A guideline for application of the Langelier Saturation Index LSI in membrane


desalination technology is provided by ASTM standard D3739.23 This sets out the
procedure for calculating the LSI for concentrates as they arise in RO desalination up
to a maximum TDS of 10,000 mg/l. Calculation of the concentration values in the
RO concentrate corresponds to the calculation procedure of ASTM D4582. For
ASTM Standard D3739, too, calculation of the saturation pH as well as the effective
pH in the concentrate and thus also determination of the concentrate’s LSI are based
on the total alkalinity mC,M,AC,T , that is, on all acid-consuming alkaline constituents.
When calculating the saturation pH, the concentration values pCa and pAlk required
for this purpose as well as the value C, this being an operand corresponding to the
activity coefficient factor fY and combines the influence of all activity coefficients,
are determined graphically by means of a nomogram. ASTM D3739 gives no details
of and references to the basis for calculating the thermodynamic parameters nor
those for the activity coefficients.
A numerical and more precise approach for calculating the SITh and LSI for water
with low to moderate salinity is provided on the basis of the standard method APHA
2330 of the Standard Methods for the Examination of Water.24 Calculation of the
thermodynamic parameters and ion products of water are based on [117, 118],
respectively. For calculating the activity coefficient, the APHA/AWWA standard
method makes use of the Davis equation (see Sect. 3.2.3.2.1, Eq. 3.95), while stating
its validity, and thus also that of the guideline with up to an ionic strength of I < 0.5.
An alternative method for calculating the calcite solubility is presented in DIN
38404-10—Calculation of the calcite saturation of water.25 This differs, though,
from APHA Standard 2330 in how the thermodynamic parameters and the activity
coefficients are calculated, which is by taking an extended Debye-Hückel equation
as calculation algorithm (see Sect. 3.2.3.2.1, Eqs. 3.94 and 3.94a). Additionally, for
this also the formation of ion pairs is incorporated into the determination of the
saturation index SI (see Sects. 3.2.3 and 3.2.3.1).

Specific Interaction Algorithms: PHREEQC and the Pitzer Database


With increasing TDS of a solution, the interactions between its constituents intensify
and, among other things, ion association components are created with charges that
differ in comparison with the mostly ionogenic status in which the solution
components are present at lower salinities. For thermodynamic modelling of the
scaling potential in seawater and in the concentrates of seawater desalination
systems, in order to be able to take the intensified interactions between the solution
components into account, variants and extensions of the equations of the Debye-
Hückel model are needed, referred to as specific interaction algorithms. Such an

23
ASTM D3739 Standard Practice for Calculation and Adjustment of the Langelier Saturation
Index for Re-verse Osmosis.
24
APHA 2330 Calcium Carbonate Saturation.
25
DIN 38404–10—Physical and physico-chemical parameters (Group C)—Part 10: Calculation of
the calcite saturation of water (C10).
5.3 Fouling and Scaling in RO Systems 567

expanded algorithm is the Pitzer model for which, in addition to the interactions
between strong electrolytes with dissimilar charges, also considered are mutual
interactions between similarly charged ions, ions with opposite charges, and
components with no charge, as these are created by ionic association.
The equations of the Pitzer Model are listed under Sects. 3.2.3 and 3.2.3.2.1,
Eqs. 3.103–3.103j and the model itself is described there in detail. However, this
model’s procedures are so complex that manual calculation is scarcely feasible. The
PHREEQC software tool that was developed on behalf of the U.S. Geological
Survey (USGS) contains various data packages for calculating the composition
and speciation, i.e. formation of component species, in aqueous solutions. Among
others, it includes the data package pitzer.dat with which speciation can be
simulated, and from this, the saturation concentrations of the solution components
and scaling potential of seawater and the concentrates of seawater desalination can
be calculated on the basis of the Pitzer model. Details of the software for thermody-
namic modelling as well as the PHREEQC computer program, its components, and
features can be found under Sect. 3.2.3.2.1.
Annex 5.A4 of this chapter contains a PHREEQC Pitzer calculation log for
seawater with a TDS of approx. 36,000 mg/l at a temperature of 25  C with the
speciation output for this water, i.e.with the concentrations of its components. The
program calculates the corresponding pH from the carbonate alkalinity mAC and the
total carbon concentration mCT of the seawater as well as the distribution of the
individual ions and ion species, and from this, the values of the saturation index of
potential scaling components. This takes into consideration both the solubility
relationships of binary compounds comprising individual cation/anion pairs and
those of mixed components for which multiple different cations and anions are
present in a compound.
In the case of the investigated seawater, it can be seen that referred to calcite
CaCO3 with a saturation index of 0.26, this is already supersaturated. The mixed
components dolomite CaMg(CO3)2 and huntite CaMg3(CO3)4 with SIs of, respec-
tively, 1.36 and 0.41 exhibit a still higher carbonate scaling potential than is the case
for the CaCO3—modification of calcite. In addition, there is a supersaturation of
magnesite MgCO3, whose saturation index of 0.35 here also exceeds the SI of
calcite.
If during desalination, the seawater of Annex 5.A4 were to be concentrated at a
recovery rate of 45% (Y ¼ 0.45) and a average specific flux of 12 l/m2, h, it will attain
the concentrate value as shown in the PHREEQC-Pitzer calculation printout of
Annex 5.A5. For the pH of 7.55 as calculated by PHREEQC from the concentrations
of bicarbonate mC,M,HCO3 , carbonate mC,M,CO2 3
, and carbon dioxide mC,M,CO2 , i.e. the
carbon alkalinity mC,M,AC , it can be seen that the above-named mixed compounds are
what determines the scaling potential of the concentrate and no longer just the
saturation index of the CaCO3—modification of calcite, as is assumed for stoichio-
metric modelling of calcium carbonate scaling.
Likewise for the concentrate from the second pass of an RO seawater desalination
plant operated with a raised pH, its scaling potential is determined by composite
568 5 Reverse Osmosis Membrane System: Core Process of SWRO

compounds made up of several components on the cation side. As shown by the


PHREEQC calculation printout in Annex 5.A6 for such a concentrate with a pH of
10.0, under these conditions dolomite CaMg (CO3)2 with a saturation index of 0.98
to 1.35 is the compound that determines the degree of scaling, followed by magne-
site MgCO3 with an SI of 0.61.
In this case, the calculations for the concentrate of the second pass were not
performed using the Pitzer database pitzer.dat, but instead the database watec4f.dat
was taken, this being likewise included in PHREEQC. The activity coefficients are
calculated in watec4f.dat using the Truesdell-Jones equation, also known as the
WATEC-Debye-Hückel equation. This means that this modelling approach is suit-
able particularly for fresh water up to the brackish water range (see Sect. 3.2.3.2.1,
Table 3.10 and Eq. 3.96). To determine the speciation, i.e. the distribution of the
solution components, for this PHREEQC calculation procedure additionally, the ion
association algorithms find application (see Sect. 3.2.3, Eqs. 3.67–3.75).
With PHREEQC, it is also possible to calculate the influence of elevated pressure
on the scaling behaviour of the solution components. Upon entering the keyword
data block “REACTION_ PRESSURE” into the program’s input file and defining a
certain pressure or pressure range to be calculated, as shown in the calculation
printout in Annex 5.A5, the solution parameters, component distributions, and
values for the saturation index for various pressure conditions can be calculated. It
then becomes apparent that with increasing pressure, the scaling potential of the low
solubility solution components decreases.
For the example in Annex 5.A5 with a reaction pressure of 75 bar, upon
comparing the values for the scalants at atmospheric pressure and those at elevated
reaction pressure, the saturation index SI reduces in part only relatively slightly in
the first and second decimal places. The change in the scaling behaviour of the
scalants is then appreciably greater for this parameter when these are referred to the
scaling ratio SR in accordance with Eqs. (5.178b) and (5.192).

SR ¼ 10SI : ð5:192Þ

Thus, according to the example in Annex 5.A5 for calcite, when comparing the
values at 75 bar with those at atmospheric pressure, the saturation index SI decreases
by approx. 12%, whereas the saturation ratio SR drops by almost 15%. For magne-
site, the SR decreases by approx. 15%, for dolomite 26%, and for huntite even by
some 45%. Thus, when modelling the scaling potential during seawater desalination,
the operating pressure of the RO system has also to be considered for the scaling
calculations.

5.3.4.2 Scaling Control and Prevention


The potential for build-up of concentration, i.e. the maximum possible concentration
factor CFmax in a desalination system, depends on which of the constituents of the
water to be treated will be the first, for the respective feed concentrations and
solubility products, to attain its solubility limit or its saturation threshold under the
prevailing operating conditions of temperature, TDS/ionic strength, and pH.
5.3 Fouling and Scaling in RO Systems 569

Therefore, in order to establish whether, with the design of the RO desalination


system, the desired recovery rate YRO can be attained for the makeup of the water to
be treated, first it is necessary to prepare a profile of the degree of saturation or
supersaturation of the water’s constituents, as is shown in Annex 5.A4. This then
shows, for the water composition, which of its low solubility constituents fixes the
maximum permissible concentration factor CFmax, i.e. the maximum possible recov-
ery rate YRO,max.
The concentration factor CFX 1 ,X 2 of a compound is calculated from its ion product
ICPX 1, X 2 ,F in the feed and ICPX 1, X 2 ,C,M in the concentrate together with the concen-
tration polarization factor βE,t in the system’s final element in accordance with
Eq. (5.193). This equation is derived from Eqs. (5.36), (5.168), (5.168a), and
(5.169e) with the simplification that the concentration factors of the two components
X1 and X2 are set equal, i.e. also their salt passage may be assumed to be the same
(CFX 1 ¼ CFX 2 ; SPX 1 ¼ SPX 2 ).
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n1 þn2 ICPX 1, X 2 ,C,M
ICPX 1, X 2 ,F
CFX 1 ,X 2 ¼ : ð5:193Þ
βE,t

For the compounds X1 and X2, the maximum concentration factor CFX 1 ,X 2 , max that
is permissible on the basis of their solubility is then calculated with, instead of using
the ion product of the compound ICPX 1, X 2 ,C,M in the concentrate, the value of its
solubility product K sp,c,X 1 ,X 2 at that location under the corresponding design and
operating conditions, as shown in Eq. (5.193a).
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n1 þn2 K sp,c,X ,X
1 2
ICPX 1, X 2 ,F
CFX 1 ,X 2 , max ¼ : ð5:193aÞ
βE,t

CFX 1 ,X 2 , max ¼ concentration factor, max. possible []


K sp,c,X 1 ,X 2 ¼ Solubility product of component X1, X2 in concentrate
The lowest value for CFX 1 ,X 2 , max of the profile is then the maximum concentration
factor CFmax that is attainable under these conditions.
The RO system’s associated maximum recovery rate is calculated again by
setting the salt passages of the two components X1 and X2 to be equal (SPX 1 ¼
SPX 2 ¼ SPX 1 ,X 2 ) in accordance with Eq. (5.194).
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ICPX X ,F
1  n1 þn2 K 1, 2  βE,t
sp,c,X 1 ,X 2
Y RO, max ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð5:194Þ
n1 þn2 ICPX 1, X 2 ,F
1 K  βE,t  SPX 1 ,X 2
sp,c,X 1 ,X 2

YRO,max ¼ recovery coefficient RO, max. possible []


570 5 Reverse Osmosis Membrane System: Core Process of SWRO

If the salt passage of the two components is neglected, which is permissible for an
RO system and for the alkaline earth sulphates CaSO4, BaSO4, and SrSO4 due to the
high rejection rates of the seawater membranes for these components, Eq. (5.194)
simplifies to Eq. (5.194a).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ICPX 1, X 2 ,F
n1 þn2
Y RO, max ffi 1   βE,t : ð5:194aÞ
K sp,c,X 1 ,X 2

For compounds with binary components like the alkaline earth sulphate scalants
with n1 + n2 ¼ 2, YRO,max is then calculated using Eq. (5.194b) and for calcium
fluoride CaF2 this is simplified with n1 + n2 ¼ 3 to Eq. (5.194c).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ICPX 1, X 2 ,F
Y RO, max ffi1  βE,t : ð5:194bÞ
K sp,c,X 1 ,X 2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3 ICPX 1, X 2 ,F
Y RO, max ffi1  βE,t : ð5:194cÞ
K sp,c,X 1 ,X 2

As shown by the profile of the degrees of saturation of its components according


to the example of Annex 5.A4, even without their build-up of concentration by the
desalination process, some of the constituents of seawater are supersaturated, like
calcite, aragonite, and other mixed alkaline earth carbonate compounds as well as
BaSO4. Therefore, in order to attain a recovery rate YRO within a range of 35% to
50%, this being the target for seawater desalination, it is necessary either:

• to correspondingly reduce the concentrations of the cations and anions of rele-


vance for scaling

or

• prevent precipitation of the scalants within the RO system.

Ion exchange or precipitation processes for softening or decarbonization as some-


times used as a pretreatment stage for brackish water desalination to modify the
concentrations of the scalants’ cations and anions have not found application for
seawater desalination up to now. What does come into consideration in this case for
cutting carbonate concentration is in particular acid dosing. By lowering the pH, this
accordingly mnimizes the proportion of carbonate ions in the carbonic acid equilib-
rium and thus also the scaling potential of the alkaline earth carbonate compounds
(see Sect. 5.3.4.2.2).
Although dosing of so-called antiscalants does not completely prevent generation
of scaling products, but despite existing supersaturation of the scalants, it can be
5.3 Fouling and Scaling in RO Systems 571

delayed to such an extent that the formation and precipitation of scaling products in
the desalination system are greatly reduced or even prevented.
For seawater desalination, mostly both conditioning processes—dosing of
antiscalant + acid – are employed in combination. When operating a seawater RO
plant in the alkaline mode, scaling is prevented solely by dosing of antiscalant. This
mode of operation has advantages if in the SWRO’s product water a low concentra-
tion of boron must be maintained (see Sect. 5.1.5.2.4).

5.3.4.2.1 Kinetics of Scaling: Antiscalants Dosing


Kinetics of Scaling
With stoichiometric or thermodynamic modelling of the solubility relationships of
the constituents of the seawater and of the concentrate resulting from desalination,
the probability of the occurrence of supersaturation is determined. Whether and the
extent to which scaling arises from the supersaturation of a certain component,
though, depends on temporal, i.e. kinetic, processes which, in turn, are functions
of specific solution parameters and operating conditions.
The formation of scaling products from a supersaturation situation usually pro-
ceeds in several phases, as shown in Fig. 5.60. First nuclei are formed that grow and
at a certain point result in the formation of microcrystals. These microcrystalline
structures grow to larger structures that are then precipitated and may form coatings
of scale. The timeframe for nucleation is termed the formation time, τn, which is
followed by the nuclei growth time, τg, with both together up to the start of crystal
formation termed the induction time, τind (Eq. 5.195).

τind ¼ τn þ τg : ð5:195Þ

τind ¼ induction time [s], [min]


τn ¼ nucleation formation time [s], [min]
τg ¼ nuclei growth/crystal formation time [s], [min]

Fig. 5.60 Phases of scaling development


572 5 Reverse Osmosis Membrane System: Core Process of SWRO

The length of the induction time τind is the decisive factor for the probability of
scale formation in a technical system. If for RO desalination the retention time τret,C
of the RO concentrate from the place where supersaturation arises to when it exits the
RO system is less than the induction time τind, there will be no scale formation during
desalination.
The induction time of a specific scalant can be determined experimentally by
measuring τind, while varying the temperature in K and the saturation ratio SR or the
saturation index SI. A logarithmic plot of τind as a function of these two parameters
then results, in accordance with Eq. (5.196), in a straight line with quotients A and B,
that are both specific for the investigated compound and how nucleation
proceeds [121].

B B
log τind ¼ A¼ 3  A: ð5:196Þ
T  ð log SRÞ
3 2
T  SI2

The quotient B is made up of various parameters that are specific for the
investigated compound and the nature of the nucleation process in accordance
with Eq. (5.196a).

β  ϑ2  γ 3S  f ðθÞ
B¼ : ð5:196aÞ
ν2  ð2:3  kb Þ3

β ¼ geometric shape factor


ϑ ¼ molecular volume of scalant [m3/mole] ¼ 3.7  105 for calcite
¼5.21  105 for barite
γ s ¼ surface energy [J/m2]
f(θ) ¼ factor differentiating heterogeneous and homogenous nucleation
f(θ) ¼ 1 for homogenous nucleation
f(θ) ¼ <1 to 0.01 for heterogeneous nucleation
υ ¼ number of ions into which a molecule dissociates ¼ sum of stoichiometric
numbers ni
kb ¼ Boltzmann constant ¼ 1.38064852  1023 [J/K]
The parameters of the quotient B that materially influence the induction time are
the surface energy γ s and the factor f(θ) that defines how nucleation proceeds. The
greater the surface energy γ s of the emerging nuclei and crystal particles, the longer
is the induction time τind. The factor f(θ) has a value of unity if nucleation proceeds
homogeneously. This means that the nuclei are formed within a solution in which
there are no foreign particles and where the developing nuclei are not in contact with
a solid surface. With heterogeneous nucleation, there is surface contact and the
nuclei are influenced in their development by the presence of solid particles.
Heterogeneous nucleation predominates in practice during nuclei formation and
crystalline growth for scale formation in membrane desalination. The lower the
value of the factor f(θ), i.e. the closer nucleation is to heterogeneous, the shorter is
the induction time τind. For scaling in practice, this means that the scale formation is
greatly influenced by the extent to which particles are swept into the flow channels of
5.3 Fouling and Scaling in RO Systems 573

the membrane elements and the deposits that are already there. Not only does the
CEOP effect thus increase due to existing deposits the scaling potential, but also
scaling is accelerated in the presence of membrane coatings in accordance with
Eq. (5.196a).
For the principal scalants in seawater, i.e. the alkaline earth sulphates and calcium
carbonate, a number of lab investigations to determine how the induction time
depends in particular on temperature and the degree of supersaturation (SR and SI)
have been published. The result of one of these investigations for calcite (CaCO3) is
shown graphically in Fig. 5.61a. Equation (5.197) shows the algorithm used for
plotting the graphs together with the corresponding coefficients for calcite CaCO3
and barite BaSO4 [122, 123].

α1 α2 α
log τind ¼ α0 þ þ þ 3 : ð5:197Þ
SI T S  T

Mineral α0 α1 α2 α3
Calcite 4.22 13.8 1876.4 6259.6
Barite 1.83 12.1 885.8 5460.3

The graphs in Fig. 5.61a show the dependency of the induction time τind of
calcium carbonate on the temperature t within a range of 1.0–2.0 of the saturation
index, SI. The SI itself, too, substantially impacts the induction time and its expo-
nential relationship with increasing SI is even greater than the influence of tempera-
ture. Above a supersaturation value of 2.5 for the SI, there is a sharp fall in the
influence of temperature and the induction time is then determined predominantly by
the saturation index.
Also as the ionic strength I increases, the induction time for CaCO3 decreases, as
shown by the curves of Fig. 5.61b. The graph in Fig. 5.61b has been calculated on
the basis of Eq. (5.198) for a temperature range of 20–25  C and is obtained by
regression analysis of test results in [121].

log τind ¼ k  n  log SR ¼ k  n  SI: ð5:198Þ

Ionic strength
mole/l k n
0.05 4.614203 1.84406
1.12 4.179697 1.545001
1.34 3.889555 1.379545
574 5 Reverse Osmosis Membrane System: Core Process of SWRO

Induction time τ of
CaCO3 [min]

100,000.0
Temperature [°C]

10,000.0 10

20
1,000.0 30
40
50
100.0

10.0

1.0

0.1

0.0
0.5 1.0 1.5 2.0 2.5 3.0 3.5
Saturation index SI

Induction time τ of
CaCO3 [min]
700
Ionic strength [mol/l]

600
0.05

500

1.12
400
Temperature = 20 - 25°C

1.34
300

200

100

0
0.5 1.0 1.5 2.0 2.5
Saturation index SI

Fig. 5.61 (a) Induction time of calcium carbonate (calcite)—dependence on saturation index SI
and temperature t. (b) Induction time of calcium carbonate (calcite)—dependence on saturation
index SI and ionic strength I

Above an SI of 1.5, the induction time is governed solely by the saturation index
and no longer by the ionic strength.
The lab tests for the investigations of [122, 123] (Eq. 5.197) and [121] (Eq. 5.198)
were conducted in both cases for homogeneous nucleation which means that the
5.3 Fouling and Scaling in RO Systems 575

results obtained for the induction time are not automatically applicable to practical
operation of desalination systems with heterogeneous nucleation. Further, these test
results were obtained for solutions of pure scalant components with the addition of
NaCl for adjustment to the desired ionic strength, so the influence of foreign ions is
not taken into account. The graphs and values of τind in Fig. 5.61a, b are therefore to
be understood as being indicative only and to show the differing dependencies of the
induction time of CaCO3 on temperature and the degree of supersaturation as well as
the influence of ionic strength. Values derived from these graphs should not be used
in actual practice for the design of RO systems.
The dependencies of the above influencing parameters may differ appreciably for
other scalants present in seawater from those shown for CaCO3.
The dependency of the induction time τind, 25  C at 25  C for barium sulphate on
the degree of supersaturation and at the temperature T in accordance with [124, 125]
is shown by the algorithms of Eq. (5.199).
" #2
1 1
¼ 0:182 þ 0:114  ln  2 : ð5:199Þ
log τind,25 C log SSR2Th

1 1000
¼  A3 þ A1 :
log τind,T T
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
γ Ba2þ  mBa2þ  γ SO2  mSO2
SSRTh ¼ 4 4

K sp,BaSO4,25 C
0

SSR A3 A1
5 2.31 4.41
10 2.16 5.23
15 1.66 4.05
20 1.34 3.28
SSRTh ¼ supersaturation ratio with thermodynamic modelling []
K 0sp,BaSO4,25 C ¼ thermodynamic solubility product of BaSO4 at 25  C [(mol/l)2]

With Eq. (5.199a) the dependency of the induction time for calcium sulphate on
the temperature [K], the saturation index SI and the ionic strength I [mol/l] can be
calculated in accordance with [126].

0:221 2171:2 1:715


log τind,0 ¼ 6:297  þ þ pffiffi : ð5:199aÞ
SI T  SI0:285 1 þ I
Calcium carbonate is the predominant scalant in RO seawater desalination.
Alongside the influence of the nature of the nucleation process, i.e. homogeneous
or heterogeneous, on the kinetics of scaling during the formation of the solid CaCO3
products, the presence of foreign ions may also strongly influence the crystal
576 5 Reverse Osmosis Membrane System: Core Process of SWRO

structures of the resulting calcium carbonate—calcite, aragonite, and vaterite—as


well as the induction time τind. Thus, investigations under RO concentrate conditions
have shown that the presence of magnesium and sulphate changes how nucleation
proceeds as well as the nature of the so-created crystal structures in that crystalliza-
tion is retarded, and in comparison with the precipitation of calcium carbonate
without these foreign ions, the induction time is appreciably prolonged [127].
In line with these test results, it would be possible under specified criteria for the
conditioning measures in RO seawater desalination systems to reduce the dosing rate
of acid and/or antiscalant or even to dispense with chemical dosing altogether.
However, such lab findings obtained under conditions of predominantly homoge-
neous nucleation would have to be tested and confirmed by trial runs of desalination
plants under conditions encountered in actual practice, in particular the presence of
particles and coatings in the elements’ flow channels as well as with heterogeneous
nucleation and crystallization. Naturally also to be considered is the scaling
behaviour of other seawater scalants like barium sulphate BaSO4 [125], strontium
sulphate SrSO4, and in particular mutual influencing of CaCO3 and CaSO4 in their
scaling kinetics [128].

Antiscalant: Mechanism and Dosing


Antiscalants, which are also termed scaling or threshold inhibitors, are polymeric
chemical compounds which when dosed to solutions enable the induction time and
timeframe for the formation of precipitation products from supersaturated scale-
forming components to be so influenced that over a certain time interval no scalant
precipitation will take place.
The first antiscalant products contained inorganic polymeric phosphates as active
compounds, like sodium hexametaphosphate, SHMP. This has an inhibiting action
on both calcium sulphate and calcium carbonate. However, if the antiscalant solution
is stored for lengthy periods and also at elevated temperatures, SHMP hydrolyses
with the formation of orthophosphates that possess no inhibiting action, but instead
form low solubility calcium phosphate with any calcium ions present. Thus, dosing
of such solutions does not produce the desired inhibiting action, but may rather result
in scaling. Further, the presence of phosphate in the concentrate of an RO system
may promote eutrophication in the seawater at its location of discharge. If the
phosphonic acid group is present in the form of an organic bond, there is virtually
no further splitting of the polymers by hydrolysis. The result of the creation of such
stable organic polymers in the form of phosphonates was to greatly enhance the
inhibitory effect due to the greater efficacy of these products. Other organic
polymers, too, like polycarboxylic acids, e.g. polyacrylic acid, polymethacrylic
acid, polymaleic acid and their salts and derivatives on their own or when mixed
with polyphosphonates or as blended polymerisates of these types of compounds,
find application as antiscalants. In part, these compounds act as dispersants or the
formulations for antiscalants and additionally contain such substances in order to
prevent agglomeration of the generating solids and their attachment to surfaces.
The molecular size or molecular weight of an antiscalant greatly affects its
effectiveness in inhibiting certain scalants. The molecular weights of polymeric
5.3 Fouling and Scaling in RO Systems 577

organic antiscalants range from 300 to 1000 Da, such as the polyphosphonates, up to
the several thousand daltons of polycarboxylic acids or of compounds formed by
copolymerization.
Antiscalants’ action mechanisms are not yet fully understood. Most explanatory
models assume that the inhibitor molecules are absorbed on the surfaces of the nuclei
as they are created and are then subsequently bound into the structure of the micro
crystals as they form. During surface adsorption, the microparticles’ charge potential
increases and at the same time also their surface energy, thus hampering and
delaying the agglomeration of the nuclei. The embedment of inhibitors in the
micro crystals as they are created prevents regular crystal structures from develop-
ing. The strong influence of molecular size on the inhibitors’ effectiveness can be
explained in that this is greater the more the nuclei’s surfaces are covered by the
inhibitor and the more it is embedded in the crystal structure.
The extent to which antiscalants are effective as inhibitors differs, i.e. the maxi-
mum supersaturation that can be attained by the various seawater scalants as
expressed by the saturation ratio SR, the saturation index SI or LSI, and the Stiff
& Davis Saturation Index S&DSI. The manufacturers of antiscalants and also the
membrane manufacturers quote upper limits for the potential degree of inhibition of
the scalants concerned. In Table 5.21, ranges of limit—so-called threshold—values
are compiled for the supersaturation parameters referred to from the datasheets of the
various antiscalant and membrane manufacturers for each seawater scalant, with and
without dosing of antiscalants. The maximum supersaturation values—SR, LSI and
S&DSI—listed in Table 5.21 for individual scalants with and without antiscalant

Table 5.21 Possible range of saturation parameters SR, LSI, and S&DSI of scalants with and
without antiscalant dosing
Range and maximum values of
Compounds Type of antiscalant SR LSI S&DSI
CaSO4, BaSO4, No antiscalant 0.8 to
SrSO4, CaF2 0.9
CaCO3 0.1 to 0.1 to
0.2 0.2
CaSO4 Organic antiscalant 2.0 to
4.0
Sodium 1.5
hexametaphosphate
BaSO4 Organic antiscalant 40 to
100
SrSO4 10 to
50
CaF2 100 to
500
CaCO3 Organic antiscalant 1.5 to 1.0 to .
.2.8 1.5
Sodium 0.5 to 0.5
hexametaphosphate 1.0
578 5 Reverse Osmosis Membrane System: Core Process of SWRO

dosing determine the peak values of the concentration factor CFmax and the recovery
rate YRO,max for the scalant concerned during desalination.
These peak values may be calculated for binary scalants like the alkaline earth
sulphates with Eq. (5.200) for CFmax and Eq. (5.200a) for YRO,max. By neglecting the
salt passage of the scalant components, YRO,max is calculated using Eq. (5.200b).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n1 þn2 SRK sp,c,X ,X
1 2
ICPF
CFmax ¼ : ð5:200Þ
βE,t
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1  n1 þn2 SRKICPF  βE,t
Y RO, max ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 2 : ð5:200aÞ
sp,c,X ,X

1  n1 þn2 SRKICPF  βE,t  SPX 1 ,X 2


sp,c,X 1 ,X 2

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n1 þn2 ICPF
Y RO, max ¼1  βE,t : ð5:200bÞ
SR  K sp,c,X 1 ,X 2

The value of the target recovery rate YRO when designing and operating an RO
desalination plant has to be selected such that it does not exceed the lowest value of
the peak recovery rates YRO,max calculated for each of the scalants.
In order to attain the maximum threshold values of each of the scalants, however,
the antiscalant of the various manufacturers of these products has to be selected that
is most suitable for the scalant profile of the RO concentrate that is to be stabilized,
and this must be injected into the feed of the RO desalination system at an
appropriate dosing rate.
An antiscalant’s effectiveness in inhibiting a certain scalant and the inhibitor’s
dosing rate needed for this purpose have to be determined empirically in lab jar tests
under the closest possible simulation of conditions met with in practice, or on pilot
plant scale, or on an industrial scale during trial operation of a desalination plant. As
investigations in accordance with [122, 123] show, this depends on the chemical
structure of the scaling inhibitor itself, the degree of supersaturation of the scalant,
and the concentration relationships in the solution as well as its pH and temperature.
In [122, 123] it is shown how, with an appropriate calculation algorithm, the
antiscalant efficiency of an inhibitor can be determined from a jar test for a specific
scalant and then, on the basis of this result, an estimate may be made of the
antiscalant dosing rate.
First, the antiscalant efficiency beff is calculated using the coefficient βi deter-
mined in lab tests for a certain inhibitor and the desired scalant with Eq. (5.201). For
this purpose, the scalant’s supersaturation, i.e. its saturation index, the temperature in
K, and the pH, must be known. The factor Rc is calculated from the ratio of the molar
concentrations of the cations and anions that determine the scaling behaviour of the
scalant. Listed in the table belonging to Eq. (5.201) are the bandwidths of βi values
for scaling inhibitors for the scalants calcium carbonate and barium sulphate. For
particular antiscalants, individual βi values are listed in [122].
5.3 Fouling and Scaling in RO Systems 579

β2
log beff ¼ β0 þ β1  SI þ þ β3  pH þ β4  Rc : ð5:201Þ
T
m m
Rc ¼ mHCO
Ca2þ

for CaCO3 ¼ m Ba2þ
2
for BaSO4
3 SO
4

Range of β coefficients
Mineral β0 β1 β2 β3 β4
Calcite 1.19 to 1.22 to +1082 to 0 to +0.2 +0.13 to
4.63 1.69 +1919 +0.29
Barite 3.23 to +0.78 1.05 to +640 to +1255 0 to +0.07 to
1.76 +0.34 +0.23
beff ¼ antiscalant efficiency [l/mg]
Rc ¼ ratio of molar concentration of components of scalant [mol/mol]

From the bandwidths of the βi values in the table associated with Eq. (5.201), it
can be seen how greatly the various antiscalant products differ in their effectiveness
in inhibiting calcite and barite scaling.
Then with the antiscalant efficiency beff so obtained, the dosing rate cantisc is
calculated for the antiscalant with Eq. (5.201a). Also a determining factor for the
dosing rate in addition to beff is the logarithmic relationship of the induction time τind,
antisc with antiscalant that is intended for inhibition of scaling and the induction time
without antiscalant that becomes established depending on operating conditions.
τind, 0 is calculated for the various seawater scalants as shown in Eqs. (5.197),
(5.198), (5.199), and (5.199a). The quotient of the two induction times is multiplied
by a safety factor fs to account for operating criteria that cannot be simulated in lab
testing.

1 τind, antisc  f s
cantisc ¼  log : ð5:201aÞ
beff τind,0

cantisc ¼ antiscalant concentration [mg/l]


fs ¼ safety factor []
τind, antisc ¼ induction time with antiscalant [s]
τind, 0 ¼ induction time without antiscalant [s]
As shown by Eq. (5.201a), although antiscalant inhibition delays the formation of
scaling products, scaling is not actually prevented. Therefore, concentrate whose
supersaturation has been temporarily stabilized in this way must be removed from
the RO system if its residence time is significantly prolonged due to a plant outage,
otherwise there is a danger of increased scaling during this time. This can be done by
flushing with the water to be desalinated at low pressure at a correspondingly
reduced recovery rate or by flushing with permeate. At a high degree of concentrate
supersaturation, this flushing operation should be carried out either straight away or
else within a short time after the plant shutdown in accordance with the instructions
of the scaling inhibitor manufacturer.
580 5 Reverse Osmosis Membrane System: Core Process of SWRO

The manufacturers of antiscalant products (American Waters Chemicals, Avista


Technologies, BASF—Sokalan, BWA Waters Additives, Genesys, GE Water, King
Lee Technologies, Nalco, Trisep, Toray-Ropur, etc.) determine the suitability of
their products for a specific application as well as the necessary inhibitor dosing rate
using software that in some cases they make available to the users of their products
and plant designers, or they permit online access to their software.
For preparing calculations with such software, normally the following must be
entered:

• a full water analysis


• water temperature
• target recovery rate of the RO system

as well as

• the membrane manufacturer together with type and designation of his membrane,
i.e. for seawater or brackish water.

The dosing rate needed for inhibiting seawater scalants normally ranges from 0.5 to
3 mg/l of the active product, i.e. well into the substoichiometric range referred to the
actual concentrations of the principal scalants. From calculations performed with
programs provided by various manufacturers of antiscalant chemicals, for seawater
in accordance with Annex 5.A4 for a 45% recovery rate and operation in the alkaline
range, i.e. a concentrate composition obtained solely with dosing of antiscalant
corresponding to Annex 5.A5 at 25  C, a dosing rate range for the manufacturers’
inhibitors of between 1 and 3 mg/l was obtained. With rising temperature, the
necessary dosing rate increases slightly or it drops with decreasing temperature.
If the presence of iron ions in an RO system from flocculant dosing in its
pretreatment stage cannot be ruled out, this has to be taken into consideration
when selecting the antiscalant and corresponding necessary dosing rate. Antiscalants
can sequestrate iron, particularly those on a phosphonate basis. Additional binding
of iron by the inhibitor, though, means that the antiscalant has to be at a correspond-
ingly higher concentration or with a specified dosing rate a corresponding proportion
of the product is used for iron sequestration, so less of the antiscalant is available for
the threshold reaction with the scalant. So the proportion of the inhibitor for iron
sequestration may be appreciably greater than what is needed for scalant inhibition
as binding of the iron to the antiscalant is stoichiometric. The required antiscalant
dosing rate will then tend to a higher value.
Antiscalants can, though, be the cause of membrane fouling or at least promote
this. Various types of inhibitor polymers, e.g. certain polyacrylates, may form low
solubility compounds with calcium and iron, and additionally, the reaction of cation-
active coagulants with certain antiscalants may result in the formation of
precipitants. This is also possible for conversion of antiscalants with oxidizing
agents through an oxidative modification of their structure and the decay products
5.3 Fouling and Scaling in RO Systems 581

that thereby form, like phosphates, as well as by a resulting reduction of the efficacy
of their scaling inhibition action.
Because the scaling inhibitors themselves or the constituents of the antiscalant
formulations may possess surface-active properties, it is important to select these
products while bearing in mind not only their effectiveness for avoidance of scaling,
but also their compatibility with the RO membranes and their influence on their
permeability and separation characteristics.
To a very large degree, antiscalants are rejected by RO membranes and remain in
the concentrate. However, regarding the quality of these products, it must be ensured
that these contain no monomeric by-products from the inhibitors’ manufacturing
process, otherwise such low-molecular compounds could infiltrate the RO permeate.
Thus, when selecting scaling inhibitors that could be used in membrane desalination
plants for generating potable water, it must be ensured that the manufacturers
provide evidence of the suitability of their products for generating drinking water
by presenting corresponding national and/or international certificates.
Because the antiscalants are discharged together with the concentrate into the sea,
they should be readily biologically degradable and also the resulting decay products
should not present a hazard to marine fauna and flora.

5.3.4.2.2 Carbonate Scaling Prevention: Acid Dosing


The potential for calcium carbonate scaling is determined by the carbonate concen-
tration at the location in the RO system where its build-up is at a maximum. It is
possible to influence this concentration by modifying the pH. By modifying the
carbonic acid equilibrium through the addition of acid to higher shares of bicarbon-
ate and carbon dioxide, the CO32 concentration can be shifted accordingly and thus
the calcium carbonate scaling potential correspondingly reduced (for further details
see Sects. 3.2.4, 3.2.4.1, 3.2.4.4, and 5.3.4.1.3). By dosing acid at an appropriate rate
at the feed to an RO seawater osmosis plant, the Stiff & Davis Stability Index in its
concentrate can be so adjusted that in accordance with the membrane manufacturers’
specifications, it has a negative value of less than 0.1 to 0.2, which means that the
calcium carbonate scaling potential is reduced correspondingly (see also Table 5.21).
However, acid dosing only influences calcium carbonate scaling. If at the same time
the solubility limit of the alkaline earth sulphates, calcium fluoride, or the mixed
products of these scalants are exceeded, antiscalant will have to be injected addi-
tionally. This means that for seawater desalination, to suppress scaling usually both
acid and antiscalant are injected. It is then possible to also exploit the delay of
calcium carbonate scaling due to antiscalant and to adjust to positive values of the
Stiff & Davis Stability Index in the concentrate of the RO seawater plant in
accordance with Table 5.21 with a lower acid dosing rate. Through this combination
of acid and antiscalant dosing to suppress scaling, the dosing rates of these two
chemical additives can be optimized so that total chemical costs are minimized.

Calculation of Acid Dosing Rate


The acid dosing rate with either hydrochloric acid HCl or sulphuric acid H2SO4 in
the feed to an RO seawater desalination plant has to be so calculated that a defined
582 5 Reverse Osmosis Membrane System: Core Process of SWRO

value for the Stiff & Davis Stability Index in its concentrate is attained or is not
exceeded. Because in this case acid dosing takes place into an enclosed system,
i.e. any CO2 that is generated cannot escape, the carbonate alkalinity is reduced by an
amount corresponding to the additionally injected acid by the valueΔmAC,acid,F . The
amount of dissolved inorganic carbon mCT,acid,F remains the same, i.e. it corresponds
to its original value in the RO feed mCT,F , as the acid dosing just changes the shares of
the components in the calcium/carbonic acid equilibrium, but there is no loss of CO2
(see Eq. 5.202).
The value of ΔmAC,acid,RO is calculated from the carbonate alkalinity mAC,F ,RO
(Eq. 5.202) prior to acid dosing, the value of the total dissolved inorganic carbon
at the RO feed mCT,F,RO (Eq. 5.202a), and the pH,acid,RO that results from acid dosing,
as shown in Eq. (5.202b).

mAC,F,RO ¼ mHCO3 ,F,RO þ 2  mCO2


3 ,F,RO
: ð5:202Þ

mCT,F,RO ¼ mCT,acid,RO ¼ mHCO3 ,F,RO þ mCO2


3 ,F,RO
þ mCO2 ,F,RO : ð5:202aÞ
 
mCT,F,RO  K 1,F  10pH,acid,RO þ 2  K 2,F
ΔmAC,acid,RO ¼ mAC,F,RO  2p  p : ð5:202bÞ
H,acid þ K
1,F  10
10 H,acid,RO þ K
2,F

ΔmAC,acid,RO ¼ change of carbonate alkalinity in the RO feed due to acid dosing


[mol/kgs]
mAC,F,RO ¼ carbonate alkalinity in RO feed [mol/kgs]
mCT,F,RO ¼ total dissolved inorganic carbon content in RO feed [mol/kgs]
pH,acid,RO ¼ pH in RO feed after acid dosing
The reduced carbonate alkalinity mAC,acid,RO due to acid dosing as well as the
changed shares of mHCO3 ,acid,RO , mCO2
3 ,acid,RO
, and mCO2,acid,RO in the carbonic acid
equilibrium due to the pH reduction are determined with Eqs. (5.202c)–(5.202f).

mAC,acid,RO ¼ mAC,F,RO  ΔmAC,acid,RO : ð5:202cÞ

10pH,acid,RO
mHCO3 ,acid,RO ¼ mAC,acid,RO  pH,acid,RO : ð5:202dÞ
10 þ 2  K 2,F
mAC,acid,RO  mHCO3 ,acid,RO
mCO2
3 ,acid,RO
¼ : ð5:202eÞ
2
mCO2,acid,RO ¼ mCT,F,RO  mHCO3 ,acid,RO  mCO2
3 ,acid,RO
: ð5:202fÞ

mAC,acid,RO ¼ carbonate alkalinity in RO feed after acid dosing [mol/kgs]


mHCO3 ,acid,RO ¼ concentration of bicarbonate in RO feed after acid dosing
[mol/kgs]
mCO23 ,acid,RO
¼ concentration of carbonate in RO feed after acid dosing [mol/kgs]
5.3 Fouling and Scaling in RO Systems 583

mCO2,acid,RO ¼ concentration of carbon dioxide in RO feed after acid dosing


[mol/kgs]
The values of the first and second stoichiometric dissociation constant of carbonic
acid as a function of temperature and salinity, which are necessary for the calculation
of ΔmAC,acid,RO , can be calculated based on Eqs. (3.123) and (3.123a), since the
calculation is done for the seawater feed to the RO sytem (see Sects. 3.2.4 and 3.2.
4.1).However, also Eq. (5.187) together with the graphs of Fig. 5.59a, b may be used
(see Sect. 5.3.4.1.3), whose values for determining the effective pHeff,C in the
concentrate are needed in any case.
After determining the ionic strength in the RO feed and calculating the changed
shares of concentration of the components mHCO3 ,acid,RO, mCO2 3 ,acid,RO
, and mCO2,acid,RO
in the carbonic acid equilibrium following acid dosing then, on the basis of the
specified recovery rate YS of the RO system, its concentration make-up is calculated
and from this the saturation pH pHs,C , the effective pH pHeff,C , and the Stiff & Davis
Stability Index constant KS&DSI,C in the concentrate are determined. From these
parameters, the Stiff & Davis Stability Index S&DSIC of the concentrate is obtained.
Details of how this calculation proceeds with determination of the influence of the
differing salt passages of the carbonic acid equilibrium components on their
concentrations in the concentrate as well as calculation of the maximum calcium
concentration at the location of peak concentration in the system’s membrane
elements are described in detail in Table 5.20 and in Sect. 5.3.4.1.3.
If with the specified acid dosing rate, the S&DSIC fails to achieve its desired
value, this calculation will have to be repeated with a different acid dosing rate.
Thereby, in line with adjustment of the pH in the RO feed, also salt rejection at the
separation membranes of the components of the carbonic acid equilibrium will
change. As the range of starting values of the feed pH for the trial-and-error calcula-
tion procedure, pH values of between 7.0 and 7.3 should be selected. This iterative
procedure is continued until the target S&DSIC value is attained.
The end result of this calculation routine is that feed pH pH, acid, RO which results
in the target S&DSIC value in the concentrate. This feed pH corresponds to an
incremental change ΔmAC,acid,RO by which the carbonate alkalinity mAC,F,RO has to be
reduced and then with which, using Eqs. (5.203) or (5.203a), the dosing rate RD, Acid,
100 % , RO needed for the desired reduction of the calcium carbonate scaling potential
in the RO concentrate is calculated.

106  TDSM,F MWacid


RD,Acid,100%,RO ¼ ΔmAC,acid,RO   ρF  : ð5:203Þ
103 Z acid
MWacid
RD,Acid,100%RO ¼ ΔmAC,acid,RO  ρF   103 : ð5:203aÞ
Z acid
RD,Acid,100%,RO ¼ dosing rate of acid to RO feed at 100% concentration [mg/l,
g/m3]
584 5 Reverse Osmosis Membrane System: Core Process of SWRO

ΔmAC,acid,RO ¼ change of mAC,F,RO by acid dosing in the RO feed [mole/kg], [mol/kg


H20]
TDSM,F ¼ total dissolved solids content of feed, mass-based [mg/kg], [ppm]
ρF ¼ density of feed water [kg/l]
MWacid ¼ molecular weight of acid [g/mole] ¼ 98.1 for H2SO4 ¼ 36.5 for HCl
Zacid ¼ valency of acid ¼ 2 for H2SO4 ¼ 1 for HCl
This calculation procedure for adjusting the S&DSI to a target value in the RO
concentrate by acid dosing into the RO feed and determining the necessary acid
dosing rate is shown in the following Table 5.22.
Since the acid dosage rate RD,Acid,100%,RO thus obtained was calculated from the
two parameters carbonate alkalinity mAC,F,RO and total dissolved inorganic carbon
mCT,F,RO of the carbonic acid equilibrium, does this acid consumption also refer only
to the acid-consuming reactions in this equilibrium and other equilibrium reactions
in seawater which influence the acid consumption and are contained in the total
alkalinity mAT , such as the boric acid/borate equilibrium, are not included in the
dosing rate thus calculated. In order to obtain the acid dosing rate that is actually
required for the practical design of a membrane desalination system, appropriate
allowances must be added to the value calculated in this way.
ASTM D458226 likewise presents algorithms with which the acid dosing rate RD,
Acid, 100 % , RO may be calculated, this time from the total alkalinity mAT . As already
explained, as a simplifying assumption the total alkalinity cAT,F is set equal to the
bicarbonate concentration cHCO3 ,F. Derived from the ASTM equation for calculating
the quotient r from the dosing rate of the acid concerned RD, Acid, 100 % , RO together
with the total alkalinity cAT,acid,RO in mg/l CaCO3 and the carbonic acid concentration
cCO2 ,acid,RO in mg/l CO2 after acid dosing (Eq. 5.204) along with the relationship for
calculating pH from this quotient r likewise in the ASTM standard (Eq. 5.186) is the
equation for calculating the acid dosing rate RD, Acid, 100 % , RO as a function of the
total alkalinity in the plant feed cAT,F and the carbonic acid content of the feed cCO2 ,F
(Eq. 5.204a). The value of the quotient r also may be calculated from pHacid,RO that
becomes established after acid dosing in accordance with Eq. (5.204b).

cAT,acid,RO cAT,F  a  RD,Acid,100%,RO


¼r¼ : ð5:204Þ
cCO2 ,acid,RO cCO2 ,F þ b  RD,Acid,100% , RO
cAT,F  r  cCO2 ,F
RD,Acid,100%,RO ¼ : ð5:204aÞ
rbþa
pH 6:2033
acid,RO
r¼e 0:423 : ð5:204bÞ

a ¼ 1.02 for H2SO4


¼ 1.37 for HCl

26
ASTM D4582 Standard Practice for Calculation and Adjustment of the Stiff and Davis Stability
Index for Reverse Osmosis.
5.3 Fouling and Scaling in RO Systems 585

b ¼ 0.90 for H2SO4


¼ 1.21 for HCl
cAT,F ¼ total alkalinity in feed [mg/l CaCO3]
cAT,acid,RO ¼ total alkalinity after acid dosing [mg/l CaCO3]
cCO2 ,F ¼ CO2 content in feed [mg/l CO2]
cCO2 ,acid,RO ¼ CO2 content after acid dosing [mg/l CO2]
r ¼ methyl orange alkalinity to carbon dioxide ratio [mg/l CaCO3/mg/l CO2]
The values for the alkalinity cAT,acid ,RO and the CO2 content cCO2 ,acid,RO in each case
after acid dosing are then calculated with Eqs. (5.204c) and (5.204d).

cAT,acid,RO ¼ cAT,F  a  RD,Acid,100%,RO : ð5:204cÞ

cCO2 ,acid,RO ¼ cCO2 ,F þ b  RD,Acid,100%,RO : ð5:204dÞ

Using the algorithms of ASTM D4582, the calculation of the value of the Stiff
& Davis Stability Index in the concentrate and the iterative process to attain the
target S&DSIC are comparable with the mode as shown in Tables 5.20 and 5.22.
For the salt passage SPAlkalinity necessary according to Table 5.20 to calculate the
value of alkalinity in the concentrate cC,M,AC , an empirical value can be used or
calculated from the salt passages SPHCO3 and SPCO32 determined according to
Table 5.10.
If for the addition of acid needed to minimize the scaling potential of CaCO3 the
only dosing of acid is at the RO feed, the concentrations of the components of the
carbonic acid equilibrium upstream of the acid dosing point correspond to the
seawater feed values for the complete RO system (mAC,F,RO ¼ mAC,F, ; mCT,F,RO ¼
mCT,F ; mHCO3 ,F,RO ¼ mHCO3 ,F etc.). The situation is different, however, if prior to
acid dosing needed for minimizing scaling there is additional acid dosing, such as for
adjusting to the optimum pH for a flocculation stage in the pretreatment section. Then
in order to calculate the RO acid dosing rate, the basis for this is the concentration
values of the carbonic acid equilibrium resulting from acid dosing for pH adjustment
in the first dosing stage (mAC,F,RO ¼ mAC,F,acid ; mCT,F,RO ¼ mCT,F,acid ; mHCO3 ,F,RO ¼
mHCO3 ,F,acid , etc.). Then regarding the dosing operations, these depend on whether
both injections of acid are into a closed system or whether dosing for the
pretreatment section is conducted in an open system. For details of the calculation
of the acid dosing rate for adjustment to a specified pH for flocculation, see Sect.
2.3.1.1.4 in Volume 2.
For such two-stage acid dosing, the acid dosing rate RD,Acid,100%,system of the
system as a whole results from the sum of the dosing rate RD,Acid,100%,Pt for
pretreatment and the rate RD,Acid,100%,RO for minimizing scaling in the RO system.
This total dosing rate RD,Acid,100%,system for the complete RO system is calculated
using Eq. (5.205).

RD,Acid,100%,system ¼ RD,Acid,100%,Pt þ RD,Acid,100%,RO : ð5:205Þ


586 5 Reverse Osmosis Membrane System: Core Process of SWRO

RD,Acid,100%,system ¼ dosing rate of acid for system at 100% concentration [mg/l,


g/m3]
RD,Acid,100%,Pt ¼ dosing rate of acid to pretreatment at 100% concentration [mg/l,
g/m3]
RD,Acid,100%,RO ¼ dosing rate of acid to RO at 100% concentration [mg/l, g/m3]

Table 5.22 Acid dosing—sequence of calculation of Stiff & Davis Saturation Index (S&DSI)
target in RO concentrate
5.3 Fouling and Scaling in RO Systems 587

Corresponding to the dosed quantity of acid, the total salt content of the seawater
and thus also its ionic strength increases in accordance with Eqs. (5.206) and
(5.206a).

RD,Acid,100%,system  MWanion acid


ΔcAnion,acid ¼ : ð5:206Þ
MWacid
cAnion,acid ¼ canion þ ΔcAnion,acid : ð5:206aÞ

ΔcAnion,acid ¼ increase of acid anion content by acid dosing [mg/l, g/m3]


cAnion,acid ¼ acid anion content after acid dosing [mg/l, g/m3]
cAnion ¼ acid anion content before acid dosing [mg/l, g/m3]
MWanion acid ¼ molecular weight of anion of acid ¼ 96.1 for H2SO4, ¼ 35.5 for
HCl
MWacid ¼ molecular weight of acid ¼ 98.1 for H2SO4, ¼ 36.5 for HCl
When compared to the existing salt concentration, though, this increase for
seawater desalination is so small that it may be neglected for the calculation
procedure as described above. However, if the S&DSI is adjusted using sulphuric
acid, the increase in the sulphate concentration should be taken into account when
calculating the scaling potential of the alkaline earth sulphates.
If the acid dosing rate is calculated by thermodynamic modelling, the calcula-
tion method corresponds largely to the procedure for stoichiometric modelling
shown in Tables 5.20 and 5.22, i.e. for calculation of the Stiff & Davis Stability
Index. However, as described under Sect. 5.3.4.1.4, for calculation with the thermo-
dynamic model the thermodynamic solubility coefficients of the carbonic acid K 01
and K 02 are used instead of the stoichiometric parameters K 1 and K 2, and the activity
coefficients of the components of the carbonic acid equilibrium, and instead of the
S&DSI, the thermodynamic saturation index SITh is calculated in accordance with
Eqs. (5.189) and (5.189a).
For this mode of calculation, the equation for determining the increments of
carbonate alkalinity ΔmAC,acid,RO is derived from the algorithm for stoichiometric
modelling (Eq. 5.202b), but with the stoichiometric dissociation coefficients of the
carbonic acid in accordance with Eqs. (5.207) and (5.207a) replaced by the thermo-
dynamic coefficients and the H+ concentration mHþ by the H+ activity aHþ
(Eq. 5.207b). This results in Eq. (5.208) for ΔmAC,acid,RO for the thermodynamic
model.

γ CO2
K 1 ¼ K 01  : ð5:207Þ
γ Hþ  γ HCO3
γ HCO3
K 2 ¼ K 02  : ð5:207aÞ
γ Hþ  γ CO2
3
588 5 Reverse Osmosis Membrane System: Core Process of SWRO

aH þ
mHþ ¼ : ð5:207bÞ
γ Hþ

ΔmAC,acid,RO ¼ mAC,F,RO

γ CO γ CO
mCT,F,RO  K 01,F  10pH,acid,RO  γ 2 þ 2  K 02,F  γ 22
HCO CO
  3 3
:
γ CO γ CO
2pH,acid pH,acid,RO
10 þ K 1,F  10
0
 γ  þ K 2,F  γ 2
2 0 2
HCO CO
3 3

ð5:208Þ

For calculating the bicarbonate concentration after acid dosing mCO2


3 ,acid,RO
from
the carbonate alkalinity mAC , Eq. (5.189d) applies and the proportion of the other
components of the carbonic acid equilibrium in a closed system is calculated with the
same relationships as for stoichiometric modelling (Eqs. 5.202, 5.202a, 5.202c,
5.202b and 5.202f).
Like when applying the stoichiometric model, the acid dosing rate is likewise
calculated with Eqs. (5.203) or (5.203a).

5.3.4.3 Uncertainties in Scaling Prediction Modelling

5.3.4.3.1 Stoichiometric Modelling


Determination of the scaling potential of seawater and the concentrates resulting
from RO desalination by means of stoichiometric modelling on the basis of the
solubility products of the pure binary scalants—alkaline earth sulphates, calcium
fluoride and calcium carbonate—as developed by DuPont in the 70s for the design of
its PERMASEP systems that was then taken as the basis for the standards ASTM
D3739, ASTM D4582, and ASTM D4692 has since become established for design
calculations of membrane desalination systems. This likewise applies for the Stiff &
Davis Stability Index whose use is included in this calculation model as an indicator
for the calcium carbonate scaling potential during RO seawater desalination. Stoi-
chiometric modelling is used by most membrane manufacturers also in their design
software, although with adjustments that have been made based on their experience.
However, stoichiometric modelling has some potential weaknesses that must be
taken into account in its application or to be compensated by extending the calcula-
tion model.
Determination of the polynomials for calculating the scalants’ stoichiometric
solubility products is mostly done using synthetic solutions for which the influence
of the salt concentration, i.e. the ionic strength, on the scalant’s solubility is mostly
simulated with a corresponding NaCl additive. The influence of foreign ions,
though, such as that of magnesium on the alkaline earth sulphate scalants as is
increasingly the case in reverse osmosis concentrates, is not sufficiently taken into
account in such testings. By considering the degree of ionic association, this
influence of foreign ions can be factored into the stoichiometric models. As
described in greater detail under Sect. 3.2.3, already in seawater and to an even
5.3 Fouling and Scaling in RO Systems 589

greater extent in the seawater desalination concentrates, ionic pairs are forming for
which the ions of the scalants themselves and also foreign ions are linked through
bonds that have just a low or no charge and are no longer fully freely available as
charged ions. In this way, the scaling potential of a scalant is lessened as it is only the
share of freely available scalant cations and anions that determine the tendency to
form scale. The way in which the calcium carbonate scaling potential is calculated
and the results obtained are influenced by the extent to which ion association is taken
into account, similar to the way in which the saturation ratio of alkaline earth
sulphates and calcium fluoride is determined. Calcium, too, forms association
bonds with a low or no charge with anions that are present, like SO42, HCO3,
and CO32, whereby the Ca2+ concentration of relevance for calcite scaling is
reduced. Also, the potential for this type of scaling is reduced accordingly, i.e. the
value of the saturation ratio SR is lessened or a positive S&DSI decreases or is
shifted to negative values.

5.3.4.3.2 Thermodynamic Modelling


In the thermodynamic modelling of seawater and the highly saline concentrates of
seawater reverse osmosis, the activity coefficients are calculated with the specific
interaction models, in particular with the Pitzer algorithms and using the PHREEQC
calculation program of the U.S. Geological Survey (USGS). However, the Pitzer
calculation mode has now also found its way into the technical calculation of
membrane systems. The membrane manufacturer Toray, for example, has included
the Pitzer calculation in its DS2 membrane design software.
As described in detail under Sects. 3.2.3.2.1 and 5.3.4.1.4, the equations of the
Pitzer model and similar specific interaction models contain a variety of algorithms,
and within these, there are interaction parameters by means of which the mutual
influencing of the charged and uncharged solution components are taken into
account when calculating the activity coefficients and thus also when modelling
the scaling potentials. The interaction parameters, though, like the stoichiometric
dissociation coefficients and solubility coefficients, are also determined for idealized
conditions [116]. With the more complex mathematical procedures of these thermo-
dynamic models, however, results can be achieved in the scaling calculations which
are quite close to the practical values measured in desalination plants. This is shown
by comparing the results of stoichiometric modelling and of the Pitzer calculation
with data measured at desalination plants when investigating the scaling behaviour
of barium sulphate [129] and also when considering calcium carbonate
scaling [130].

5.3.4.3.3 Scaling Calculation Results with the Software Tools of Membrane


Manufacturer and Antiscalant Suppliers and the Pitzer
Thermodynamic Model
If the scaling relationships in the concentrate of an RO seawater desalination system
are calculated under comparable conditions using the software of various membrane
manufacturers and antiscalant suppliers, the result is, in part, a widely differing range
of outcomes. Table 5.23 sets out the results of calculations using such current design
590 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.23 Comparison of scaling potential and acid consumption calculation results from the
various software systems of membrane and antiscalant manufacturers
Seawater analysis
Components Unit Concentration
Sodium Na+ mg/l 11.065
Potassium K+ 409
Calcium Ca++ 423
Magnesium Mg++ 1.316
Strontium Sr++ 8.2
Barium Ba++ 0.11
Chloride Cl 19.866
Nitrate NO3 2.3
Sulphate SO4 2.781
Bromide Br 69.0
Fluoride F 1.3
Boron B 4.0
Carbonate alkalinity mAC mmol/l 1.8858  103
Total inorganic carbon mCT 1.8448  103
Operation conditions of RO system
Recovery % 45

Temperature C 25
Specific flux JP,S,Ø,s l/m2, h 13
Type of membrane – Seawater—high rejection
Range of scaling potential calculations in concentrate
Saturation ratio SR%,C
• Calcium sulphate CaSO4 % min. 39; mostly 45–48; max. 58
• Strontium sulphate SrSO4 min. 3; mostly 33–37; max. 45–
57
• Barium sulphate BaSO4 min. 680; mostly 690–940; max.
2148
• Calcium fluoride CaF2 min. 0–6; mostly 84–760; max.
1110
Stiff & Davis Stability Index S&DSIC at – min. 0.41 to 0.76; mostly
feed pH ¼ 7.3 0.09 to 0.26;
max. + 0.06 to +0.09
Feed pH at S&DSIC ¼ 0 – min. 7.20–7.24; mostly 7.40–
7.50;
max. 7.6–8.0
Range of sulfuric acid consumption calculation
Sulphuric acid consumption RD,Acid,100% mg/l H2SO4 min. 6.1; mostly 7.5–8.2; max.
for feed pH ¼ 7.3 100% 13.8–16.2
Sulphuric acid consumption RD,Acid,100% mg/l H2SO4 min. 0.8–3.1; mostly 4.4–9.0;
for S&DSIC ¼ 0 100% max. 11.4–17.6

and computation software systems for a reference case for which seawater with a
composition as listed in Table 5.23 is desalinated at a recovery rate of 45%, a
temperature of 25  C, and an average specific flux of the RO system of 13 l/m2h.
5.3 Fouling and Scaling in RO Systems 591

First, the percentage values of the saturation ratio SR as obtained as outputs for
the alkaline earth sulphates and CaF2 are compared. The percentage saturation ratio
SR% is calculated from the SR by the calculation software of the membrane and
antiscalant companies in accordance with Eq. (5.209).

SR% ¼ SR  100: ð5:209Þ

In turn, the saturation index SI is obtained from the SR% with Eq. (5.209a) and
SR% from the SI using Eq. (5.209b).

SR%
SI ¼ log : ð5:209aÞ
100

SR% ¼ 100  10SI : ð5:209bÞ

Then the values of the Stiff & Davis Stability Index S&DSIC are compared, as
they are calculated for the concentrate of the membrane desalination plant by the
various programs as a measure for the calcium carbonate scaling potential when the
feed pH is set to 7.3. Further, also the values calculated for the feed pH are compared
for which, in line with the results of the various programs, an S&DSIC of 0 is
obtained in the concentrate. For the latter of these two conditions, compared are the
necessary acid dosing rates RD,Acid,100% for 100% sulphuric acid in the RO feed that
are output by each of the programs.
Within the range designated “mostly” are the computational results that are
output by most manufacturers’ software. Listed in the ranges min. and max. are
greatly deviating values as calculated by certain programs.

Comparison of the Results of the Commercial Membrane Design and Scaling


Prediction Tools
For the scalants calcium sulphate and strontium sulphate, the values calculated by
the programs for the saturation ratio are in the “mostly” range and the spread of
values within this is relatively small. The bandwidth in the “mostly” range for
barium sulphate is wider as is also the bandwidth of its minimum and maximum
values.
Because the solubility of strontium sulphate is greatly influenced by the magne-
sium concentration [109], the inclusion of the ionic association model in the scaling
calculations of one of the companies could explain the low value of the saturation
ratio SR%,C of 3% for strontium sulphate.
Particularly marked is the influence of foreign ions and here in particular that of
magnesium on the solubility of calcium fluoride. Apart from the graph shown in
Fig. 5.55 of Sect. 5.3.4.1.3 for the solubility of calcium fluoride versus ionic
strength, there are hardly any reliable data on the solubility of this compound
592 5 Reverse Osmosis Membrane System: Core Process of SWRO

under concentrate conditions in the literature. The influence of foreign ions is not
considered in this graph. However, figures for the solubility of CaF2 in seawater are
presented in [113] and these values are significantly above those derived from the
polynomial shown in Fig. 5.55. This could be the reason why, for calcium fluoride,
the bandwidth of the calculated values regarding its scaling potential for the
investigated conditions shows a range for SR%,C from 12% to 1110%, possibly
depending on the extent to which the respective company took the influence of
foreign ions into account. Additionally, in this case also the “mostly” range has the
greatest spread, from 84% to 760%, of all scalants.
When calculating the scaling potential of calcium carbonate in the RO concen-
trate with acid dosing to a target pH of 7.3 in the RO feed, values resulting from the
manufacturers’ calculation models for the Stiff & Davis Stability Index S&DSIC are
within a range from 0.41 to +0.09. Within the “mostly” range, the bandwidth of
S&DSIC values is from 0.09 to 0.26. The maximum values calculated for
S&DSIC from +0.06 to +0.09 correspond to a positive scaling potential, whereas
the calculation results in the “mostly” range exhibit undersaturation, i.e. a tendency
for calcite dissolution.
Similarly broad is the bandwidth of the manufacturers’ results when calculating
the RO feed pH, for which the scaling potential of calcium carbonate in the concen-
trate is reduced to such an extent that the S&DSIC is zero. The corresponding pH
values of the “mostly” range are between 7.40 and 7.50, and the minimum pH values
calculated for this equilibrium condition are from 7.20 to 7.24. As a maximum, pH
values from 7.6 to 8.0 are calculated.
The differing outcomes of the companies’ software for the ensuing values of
S&DSIC and also pH under the above stated reference conditions for which the
S&DSIC in the concentrate attains equilibrium naturally also have an influence on
the bandwidth of the respective acid dosing rate, as indicated by the calculation
programs, in order to attain the specified conditions. In order to adjust the pH in the
RO feed to 7.3, the values for the acid dosing rate in the “mostly” range are from 7.5
to 8.2 mg/l for 100% H2SO4. However, several membrane and antiscalant suppliers
calculate maximum values of between 13.8 mg/l and 16.2 mg/l for 100% H2SO4. In
order to adjust to the equilibrium conditions for calcite saturation in the concentrate,
that is with a value for S&DSIC of zero, the acid dosing rate for 100% sulphuric acid
lies within the “mostly” range of 4.4 to 9.0 mg/l. Also in this case, though, some
companies state maximum values of from 11.4 mg/l to 17.6 mg/l, whereas with
software from other manufacturers, minimum values are calculated of less than
0.8 mg/l to 3.1 mg/l for 100% H2SO4.
The differences in the computation results could be attributable in part to:

• the companies’ differing values for the degree of scalant rejection by the
membranes
• the application of differing concentration polarization factors or negation of these
influences in the scaling calculations, for example in the ASTM standards
• the use of proprietary polynomials for calculating the stoichiometric solubility
products of the scalants.
5.3 Fouling and Scaling in RO Systems 593

• Differing safety factors included by the companies into the calculation of the
scaling potential and acid dosing rate

Whether the influence of the elevated operating pressure for seawater desalination
on the seawater scalants’ scaling potential is taken into account in the companies’
calculation models is not apparent in their output logs. Should this differ between
calculation models, this could also be a reason for the divergent outcomes of the
scaling potential calculation.
To account for minimum values of the SR%,C that are appreciably below the
“mostly” values of the classical stoichiometric calculation approaches, a possible
explanation is that the calculation method’s model for the formation of ionic pairs,
i.e. the ionic association, has been incorporated into the calculation of the scaling
potential.
The broad bandwidth of the calculation results from the manufacturers’ and
antiscalant suppliers’ software regarding the potential for calcium carbonate scaling
and the acid dosing rate needed to minimize scaling may be explained, at least in
part, by the number of possible calculation methods. This could also be explained by
the choice of the algorithms available in the technical/scientific literature to be used
for the determination of the stoichiometric dissociation coefficients of carbonic acid
and the stoichiometric solubility coefficient of calcium carbonate or which proprie-
tary calculation equations are finding application in the software of a certain
company.
The calculation methods and the bandwidth of the resulting values for the CaCO3
scaling potential might be determined to a large part by the following:

• the extent to which calculation of the pH in the RO feed and the concentrate is
done with the simplified relationship in accordance with Eq. (5.184d), or with
Eq. (5.186) of ASTM D4582, with the quadratic equation of Eq. (5.184c), or by
taking an iterative approach from the total alkalinity mAT , with the two dissocia-
tion coefficients of carbonic acid K*1 and K*2 and under consideration of the
stoichiometric ion product of the water K w ,
• which of the polynomials from the techno-scientific literature is used for deter-
mining the stoichiometric dissociation coefficients of the carbonic acid as a
function of temperature and the salt concentration/ionic strength of the concen-
trate, or in a company’s software in the mode of calculation proprietary algorithm
is applied.
• the extent to which the concentration polarization of the scale-forming
components, i.e. their concentration mC,M,X i at the membrane wall, is taken into
consideration instead of the concentrate concentration mC,X i
• which values for the membranes’ degree of rejection RXi for the participating
components Ca2+, HCO3, CO32, and CO2 form the basis of the calculation and
how during the iterative procedure in accordance with Table 5.22 the rejection
rates are modified.
594 5 Reverse Osmosis Membrane System: Core Process of SWRO

Results of the Thermodynamic Modelling with the Pitzer/SIT Algorithms


Regarding the scaling potential of the alkaline earth sulphates and of calcium
fluoride for the reference conditions of Table 5.23, with the Pitzer model for the
two scalants, calcium sulphate CaSO4 and strontium sulphate SrSO4, roughly the
same values are calculated for the saturation ratio SR%,C in the concentrate,
amounting to around 46%. Barite BaSO4 exhibits a high supersaturation of 513%.
For calcium fluoride, by applying the PHREEQC/SIT calculation model a value for
SR%,C of 21% is determined. Consequently, the SR%,C value for calcium sulphate is
in the “mostly” range for stoichiometric modelling, with the value for strontium
tending more to its “maximum” range. For BaSO4, with thermodynamic modelling,
a significantly lower supersaturation level is obtained than is the case with conven-
tional modelling. This applies likewise for calcium fluoride whose SR%,C value with
thermodynamic modelling tends more to the “minimum” range in the comparison
table for stoichiometric modelling.
For the thermodynamic calculation of the potential for calcium carbonate scaling
with the PHREEQC and the Pitzer database, in addition to the values of the scaling
index for the binary CaCO3 compounds calcite and aragonite, also values for
composite compounds are calculated, for example dolomite CaMg(CO3)2, that for
this model exhibit a higher scaling potential than is the case for the binary
compounds (see Annex 5.A5). Therefore, in this instance, it is no longer the degree
of calcite supersaturation that determines the potential for carbonate scaling, but that
of the multi-component composite compounds that are formed from calcium and
carbonate with foreign ions, such as magnesium. If for these compounds too, the
scaling potential is to be reduced to the SI equilibrium value of zero by acid dosing, a
higher acid dosing rate in comparison to that needed for calcite is then necessary,
i.e. there must be an even greater reduction of the pH in the RO feed.
The Toray design program DS2 makes it possible, for the conditions shown in
Table 5.23, to calculate the necessary acid dosage rate for setting the Stiff & Davis
Stability Index in the concentrate to the equilibrium value of S&DSIC ¼ 0, as well as
to calculate with the Pitzer module which acid dosage is necessary to bring the
calcite saturation index SICaCO3 and also the other supersaturations of the carbonate
scaling by composite compounds, such as dolomite by acid addition, to SI ¼ 0.
The equilibrium setting to S&DSIC ¼ 0, calculated according to the conventional
stoichiometric method, is then possible with 5.7 mg/l H2SO4 100% at an feed pH of
7.5. In order to bring the calcite saturation index into equilibrium SICaCO3 ¼ 0, the
Pitzer module calculates an acid quantity of 16.4 mg/l at an feed pH of 6.9 and in
order to convert the entire calcium carbonate scaling potential including the com-
posite compounds into the state of equilibrium by acid dosage, 32.8 mg/l H2SO4
100% must be added. This results in a feed pH of 6.5. With the PHREEQC/Pitzer
software tool, comparable acid dosing rates and feed pH values are determined.
The acid dosing rate and RO feed pH as calculated for seawater desalination by
stoichiometric modelling of the calcium carbonate scaling, that is up to setting the
value of S&DSIC to zero, therefore are not sufficient for thermodynamic modelling
to completely eliminate the calcite scaling potential and also that of the composite
5.3 Fouling and Scaling in RO Systems 595

component carbonate scaling, i.e. to attain an equilibrium SI value of zero for all
these components.

5.3.4.3.4 Selection of Type and Rate of Chemicals Dosing


The reduction of the scaling potentials for a membrane seawater desalination plant is
possible, as described above, by combined dosing of acid and antiscalant as well as
by injecting just antiscalant on its own.
When dosing acid and antiscalant together, it is possible, referred to the reference
conditions of Table 5.23, to set the desired acid dosing rate as follows:

• as recommended by the membrane manufacturers to adjust S&DSIC to zero,


i.e. in this case at approx. 6 to 7 mg/l of 100% H2SO4
• in accordance with the results of the Pitzer calculation:
• to minimize the calcite scaling potential with a rate of approx. 16 to 17 mg/l of
100% H2SO4
• to eliminate carbonate scaling altogether with a dosing rate of 33 to 35 mg/l of
100% H2SO4

In the combined dosing case by additionally dosing antiscalant, scaling of the


composite carbonate compounds together with the alkaline earth sulphates and the
calcium fluoride is controlled. With a correspondingly higher dosing rate for the
antiscalant and without acid dosing, i.e. through the alkaline mode of operation, the
potential for both types of scaling can be minimized by dosing the antiscalant on its
own. Up to now, combined dosing of acid and antiscalant has been the method most
commonly applied for scaling control. Which of these two possibilities is selected
depends mainly on the two factors of operational reliability of the RO system and the
ongoing cost of the chemicals.
The extent to which the scaling potential as calculated for an RO membrane
desalination system will result in actual precipitation depends essentially on the
kinetics of each scalant under the prevailing operating conditions (see Sect.
5.3.4.2.1). If after setting the maximum scalant concentration the residence time of
the concentrate in the RO system is less than the respective induction time τind,
i.e. the time needed for precipitation, then despite a high scaling potential there will
no formation of coatings in the system. The scaling potential is only one of the
parameters that determine the induction time τind. Alongside scalant-specific factors
and how nucleation proceeds (see Eqs. 5.196 and 5.196a), both the temperature
(Eqs. 5.197, 5.199, 5.199a and Fig. 5.61a) and the TDS, i.e. the ionic strength
(Eq. 5.198 and Fig. 5.61b), decisively influence the kinetics of a scalant and how
extensively scaling takes place under the operating conditions.
With the mathematical modelling of the scaling potential, orientation values are
obtained. However, this methodology only covers partial aspects of the prediction of
scaling. With regard to chemical dosing when designing an RO system, the
uncertainties described in the determination of the scaling potentials, and here in
particular the potential of the calcium carbonate scaling, as well as the question of
the influence of the kinetics on the probability of the occurrence of scaling, imply
596 5 Reverse Osmosis Membrane System: Core Process of SWRO

that both the results of several calculation programs of membrane manufacturers and
antiscalant suppliers, as well as their respective practical experience, should be used
for the decision regarding kind of dosage and the dosage rate for acid and/or
antiscalant. The respective calculation mode on which their recommendation is
based should be clarified with the companies in line with the criteria described
above and it should also be clarified to what extent the proposals for the method
of dosing and the dosing rate of the chemicals are based on corresponding practical
experience with seawater desalination plants in operation.

5.4 RO Membrane Process Design with Membrane


Manufacturers’ Design Software

The practical design of a reverse osmosis membrane desalination system requires


extensive and complex calculations with, in part, iterative calculations, as described
in detail in Sect. 5.2.2. Therefore, it is very costly and time-consuming, and in part
only possible to a limited extent, to manually determine and optimize the configura-
tion of the membrane elements in such a system within its arrays, stages and modules
to such a depth of detail that the specifications of the membrane manufacturers
regarding the inflow, the throughflow, the recovery rate, the permissible values of the
concentration polarization, and the differential pressure for each membrane element
at its location in the membrane system will be adhered to. Local flow and concentra-
tion conditions in the individual elements naturally also determine the composition
of the product and the required operating pressure of the membrane system, i.e. its
product flow and the make-up of its partial flows.
All membrane manufacturers therefore provide software tools for detailed calcu-
lation and optimization of membrane desalination systems with their membrane
elements. As well as facilitating the detailed membrane calculation for the planner
of such systems, such programs also incorporate the experiences the membrane
manufacturers have made with their products, so insuring correct design and
handling of the membrane elements for the envisaged design operating conditions.
Further, the software tools also include safety factors defined by the manufacturers
as they deem necessary for the design and reliable operation of the desalination
process.
The manufacturers’ software incorporates algorithms and calculation procedures
that are basically the same as those described above under Sects. 5.1.5 and 5.2.
However, certain detailed aspects of the calculation methodology are modified and
adjusted based on the experience and know-how of particular membrane
manufacturers so that, although the calculation procedures as described may be
regarded as the basis for the membrane design of an RO system, they may differ
from manufacturer to manufacturer and yield different results, as shown, for exam-
ple, in the calculation of the scaling potentials under Sect. 5.3.4.3.
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 597

5.4.1 Structure of RO Design Programs of the Membrane Suppliers

Although the membrane manufacturers’ calculation programs differ in the graphic


design of their user interfaces, they are, though, quite similar with regard to the input
parameters required for membrane design, the calculations to be performed by the
design routines, and the calculation results that are output. Figure 5.62 shows the
input and output structure of a computer program for calculating a membrane
desalination design in a single-pass configuration, as typically applied in membrane
manufacturers’ software.

5.4.1.1 Necessary Input Parameters


The input structure is broken down into parameters that:

• specify the basic design of the RO system, such as


– its desired product performance criteria like product-side capacity, composi-
tion, and recovery, as well as
– the type, composition, and quality of the intake water, the type of water
extraction, and the method of pretreatment of the RO feed
• define the system configuration, such as
– location and type of feed conditioning with acid or antiscalant
– the number of RO system passes and whether there should be a concentrate
return from the second pass to the main desalination train in a two-pass system
– whether split partial mode finds application in the main desalination train
(Sects. 5.2.3.2 and 5.2.3.2.1)
• specify the membrane configuration with the necessary inputs
– whether the hybrid configuration is to be used in the RO systems (see Sects.
5.2.3.1 and 5.2.3.1.2)
– which membrane types are to be used with their specific design criteria
– the approximate figures to be applied for calculating the element and module
configuration, such as:
number of concentrate stages per array
number of membrane elements in the system
number of membrane elements per module
number of modules in the system
number of modules per stage
– the membrane design lifetime with the resulting values for the increase in salt
passage and the decrease in the membrane elements’ permeability and
product flow.
598 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.62 Membrane manufacturer design software—Basic structure, input and output
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 599

With most of these programs, the consistency of the inlet analysis is checked
when it is entered by calculating the electrical neutrality of its components. If there is
a deviation, an adjustment is proposed and also implemented.
Based on how the feed water is extracted and its quality parameters, like SDI, MFI,
TOC, etc., as well as the pretreatment process, most manufacturers’ software systems
propose design parameters for the configuration of the RO system and the membrane
elements according to Table 5.12, such as the average specific product flux of the
system, the maximum specific flux, and the permissible recovery rate of the elements.
However, these specifications can be amended to match the designer’s own concept.
With most programs, when specifying the membrane age, the increase in salt passage
and reduction in product flow of the membrane elements at the end of the specified
membrane age are calculated and proposed for input. Also, typically these calculation
values can be confirmed for input or amended by the designer (Sect. 5.2.2.1).
The approximate values to be entered into the membrane design program for the
number of RO stages, the number of membrane elements in the system and its stages,
the number of elements in the modules, and the resulting number of modules in the
system and in the stages are calculated manually as described in Sect. 5.2.2.2. These
indicative values must then be adjusted in an iterative approach as the detailed
calculation of the desalination system proceeds until they match the values laid
down by the membrane manufacturer for his membrane elements as selected for the
design. These design specifications and guidelines of the respective manufacturers
can be found in their design software either in the design-specific part of the
program’s help function or they may be called up directly (see Table 5.12 for the
design guidelines of the membrane manufacturers).
Taking the input parameters for the basic and system designs as well as the design
of the membrane elements and modules as listed in Fig. 5.62 and commented on
above, the membrane design software then calculates the corresponding figures for
the scaling potentials of the RO system, the detailed system and membrane element
configuration as well as the associated values for the mass flow and the compositions
of the part streams.

5.4.1.2 Adjustment of Software Default and Estimated Design Data


and Parameters
Manual interventions are necessary if the program issues warnings during calcula-
tion runs that design guidelines have been violated. The limit values specified by the
membrane manufacturers for the permissible operating conditions of their membrane
elements provide the background for the calculation procedures, so the program will
issue corresponding messages if these are exceeded during the system calculation.
Depending on the parameters for which the warning was issued, the designer will
then have to amend the input values for the program so that the permissible values
are no longer exceeded.
Such warnings are displayed if:

• the maximum permissible values of the operating pressure, differential pressure,


temperature, or pH of the membrane elements selected for the design have been
exceeded (see Table 5.2)
600 5 Reverse Osmosis Membrane System: Core Process of SWRO

• the design guidelines for the average specific product flux of the RO unit or, for
the individual membrane elements, their permissible concentration polarization
factors, the recovery rates, or other limit values for their inflow or throughflow are
exceeded (see Table 5.12).

So as to attain the values specified by the membrane manufacturer, appropriate


manual adjustments of the module and membrane element configuration have to be
made during the calculation run. On the membrane element level, this could mean a
change in the number of elements per module, on the level of the stage configuration
a change in the number of modules in each stage of the array, and on the system level
a corresponding adjustment of its recovery rate. Modification of the RO system’s
configuration will have to continue by applying a trial and error approach until the
computer program no longer displays any warnings.
Also with regard to the scaling potentials of the scaling-forming compounds
present in the RO inlet, the membrane manufacturers’ software systems issue
warnings if the solubility limits of individual scalants are exceeded at the recovery
rate selected for the system design. The scaling potentials for individual scalants are
shown and it is noted for which of these dosing of acid and/or antiscalant is
necessary in order to be able to maintain the specified recovery rate. Based on the
manufacturers’ past experience of dosing antiscalant, their design software also
indicates the potential ranges of supersaturation (see Table 5.21). With the supersat-
uration values of the scalants present in seawater that are possible by conditioning
the RO feed through chemical dosing, the system recovery rates of up to a maximum
of 50% common in the design of membrane desalination plants can be reliably
achieved, as shown by calculations with Eqs. (5.194b) and (5.194c) in Sect. 5.3.4.2.

5.4.1.3 Results and Output Data of the Software Calculation


The calculation results and output values of the membrane design software can be
assigned to three data groups, namely data related to:

• the RO system
• the array and how it is configured in stages
• the modules and the membrane elements they contain

The output data referred to the RO system contain:

• the mass flow of each part stream, i.e. feed, concentrate, and product, together
with its composition and temperature and the resulting recovery rate of the RO
system as a whole
• the specified membrane design conditions
• the calculated scaling behaviour with the scaling potential of each scalant and a
proposal for the acid dosing rate.
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 601

The calculated data referred to an array contain:

• the recovery rate of the array


• the pressure conditions in its part streams, in particular the necessary feed
pressure to the array
• the feed composition as modified by chemical dosing and by possible recircula-
tion of concentrate from a second pass
• the mass flows of permeate and concentrate of the array together with their
compositions
• the array’s module and membrane element configuration.

If the RO system has only a single stage as is usually the case for the principal
desalination system of a seawater membrane plant, the data listed in Fig. 5.62 under
RO array/stage values for this group correspond to the configurations of both the
array and the single stage. If the membrane desalination array contains several
stages, though, this section of the program’s data output is extended by the output
listing for the values of the array and additionally by the listings of the values for its
stages, i.e. the program printout then contains not only the results of the overall array
calculation, but also those for the calculations for its stages in the output structure as
shown in Fig. 5.62.
The membrane element data group contains for each element:

• its recovery rate


• its pressure conditions in the feed and concentrate lines, i.e. pressure and differ-
ential pressure, as well as in the permeate line
• the mass flow, composition, and osmotic pressure of the element part streams, that
is for feed, concentrate, and permeate
• the concentrations of the solution components and the osmotic pressure at the
membrane wall of the respective element
• the specific design data of each element; that is:
– specific flux
– net driving pressure (NDP)
– concentration polarization factor.

The performance data and the compositions of the part streams, in particular those
of the products, are calculated at the data group levels of module, array, and overall
system from the values of the data group for the membrane elements (see Sect.
5.2.2.3.2). The design-specific data for the membrane configuration are also calcu-
lated, and these must be consistent with the membrane manufacturers’ specifications
(Table 5.12) as the design of the complete RO system proceeds.
With some programs, the scope of the documentation of the results of the
membrane design calculation may be selected, i.e. a detailed report with all calcula-
tion data down to the mass and pressure balances as well as the composition of the
part streams can be output at membrane element level, although it is also possible to
create a so-called overview report, which just contains the calculation results at the
level of the concentrate stages. But for this type of report, the design-relevant
parameters at the membrane element level that the designer has to know in order
602 5 Reverse Osmosis Membrane System: Core Process of SWRO

to adapt the configuration to the specifications of the membrane supplier, i.e. the
concentration polarization factors, the recovery rates of the individual membrane
elements as well as their inflow and outflow ratios, are included in the abridged
calculation documentation.
For calculating a two-pass seawater desalination system, the input parameters on
the system level of the design program must also include the information needed for
the construction of this permeate-staged configuration, i.e. the specifications for the
recovery rates of the first pass Y1 and of the second pass Y2 as well as the capacity
factor fC,RO2 for the product flow of the second pass. At the system level, the output
value for the total recovery rate of the system Y1,2,RO is then obtained from the input
values for the respective recovery rates Y1 and Y2 and the capacity factor fC,RO2 for
the second pass according to Eq. (4.58). In the case of a two-pass desalination
system, the output listing of the membrane calculation is extended by the data for
the second pass. Since this contains several concentrate stages, the data documenta-
tion becomes correspondingly more voluminous and additionally contains the cal-
culation values for the individual stages and, if desired, also the results for the
membrane elements and their configuration in each stage.
A typical overview report for the membrane design of a two-pass seawater
desalination plant that documents the calculation results as described above down
to stage level and with transfer of the design-relevant data from membrane element
level into this is contained in Annex 5.A7. In this case, the second pass of the system
is designed for the full product flow of the RO system, i.e. for a capacity factor fC,
RO2 of 1. Serving as data source for this is the design software DS2 of the membrane
supplier Toray. When calculating the scaling potentials, a distinction is made
between the potential derived from conventional stoichiometric modelling and the
so-called “Pitzer % solubility”, which results from thermodynamic modelling using
the Pitzer algorithms (see Sects. 5.3.4.3 and 5.3.4.3.2).
The pressure conditions calculated by the membrane software, such as the inlet
pressures to the membrane system and to the stages and modules as well as the
differential pressures in the membrane elements and their sum for the complete
membrane system, are normally with reference to “clean” membrane elements with
no allowances made for fouling and scaling. Likewise disregarded are all pressure
losses due to piping and fittings in the inlet lines to the arrays and between the stages
as well as those arising in the pressure vessels of the modules. In some cases,
corresponding values may be entered into the membrane software so that the output
results also take account of such influences. It is though preferable to calculate only
the membrane-relevant pressure data with the manufacturers’ software and to take
account of all additional pressure losses in separate calculations.
With all of the membrane manufacturers’ software systems, it is possible to
calculate a two-pass configuration in split partial mode (see Sect. 5.2.3.2.1). Most
software tools also allow calculation of internal staging within the membrane
modules by selecting different membrane types at the modules’ feed and discharge
ends (hybrid mode or internal staging design, see Sect. 5.2.3.1.2).
With most of the membrane design tools, it is also possible to estimate the energy
consumption of a selected system and membrane configuration. For this purpose, the
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 603

appropriate energy recovery systems are selected in the programs and, after entering
the corresponding figures like the efficiencies of the HP system pumps and of the
energy recovery devices (ERDs), the degree of energy recovery and also the
influence of the selected membrane configuration and the ERDs on the energy
consumption of the membrane system are compared and optimized. However, the
approximate energy consumption figures obtained in this way only apply to the RO
membrane system under the specified design conditions and should not be taken as
being representative for the complete SWRO system (see also [13] Chap. 8 in
Volume 2). More details on in-depth calculation of the energy recovery systems of
a membrane seawater desalination plant can be found under Sect. 5.5.2.2.3.

5.4.2 Mode of Design Calculation and Application of the Software


for Changing Seawater Feed and Operation Conditions
of the SWRO

An RO system is required to produce at the specified performance level regarding


product flow, recovery rate, and product salt concentration over the full bandwidth of
its operating conditions. These operating conditions span:

• the full conceivable range of feed salt concentration


• the range from the minimum to the maximum feed temperature
• the desalination plant’s start-up conditions, that is with new membranes or, after
their stabilization, up to the specified design age.

However, calculation of the membrane system with the manufacturer’s software


only covers the operating conditions of the plant at one operating point, i.e. for
specific values for seawater salinity, temperature, recovery rate, product salt concen-
tration, and membrane age. This means that to cover the entire spectrum of
conditions under which the membrane system is to operate, a whole set of membrane
calculations under selected operating variables must be prepared. With a two-pass
system, this modelling of the operating conditions also determines the extent to
which the product performance of the second pass has to be adapted to the RO
system’s overall performance to achieve a consistent product quality.
The design documentation of an RO desalination system thus consists of a
portfolio of operating projections with each one calculated individually. The
corresponding process flow diagrams and mass balance sheets for the individual
operating cases are then created from the documented calculation data in the course
of the RO plant design.
Section 5.2.2.4 and Table 5.14 presents a compilation of the minimum design
cases needed for constructing a membrane desalination plant. There the structured
procedure for simulating the entire operating range for calculating membrane desa-
lination systems is set out in detail. In this section, the influences of the various
operating parameters on the performance of an RO system are also described
in-depth with their dependencies shown in Figs. 5.34 and 5.35.
604 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.4.3 Membrane Design Allowances and Safety Margins

When specifying their membrane elements, the manufacturers characterize their


performance related to salt rejection and product flow by way of the NaCl rejection
rate and the product flow at the standard test conditions defined for the membrane
type (see Sect. 5.1.5.2.3).

5.4.3.1 Salt Rejection


For the separation characteristics of a membrane element, a nominal value is stated
for its salt rejection but also a minimum value below which the NaCl rejection rate is
not allowed to drop. Both values are for stabilized membrane conditions, i.e. after the
membrane system has been in operation for a time period specified by the manufac-
turer. The salt rejection rates of each of the manufacturer’s membrane elements are
within this nominal-to-minimum bandwidth. The more membrane elements an RO
unit contains, the more the system’s standard salt rejection is shifted towards the
nominal value. One manufacturer states that a good approximation to the nominal
value is achieved with around 36 membrane elements in the RO system. If an RO
unit contains far fewer membrane elements, its standard salt rejection rate may even
approach the specified minimum value.
The calculations for the design of a membrane system as performed by the
membrane manufacturers’ software are based on the nominal values of the standard
salt rejection rate. If the result of such a membrane plant performance calculation
with a manufacturer’s software tool is compared with the actually attained RO
values, the measured values for salt rejection, salt passage, and concentration of
certain components i ci, P, S(d ), eff or of the TDS in the RO product TDSP, S(d ), eff may
differ from the values calculated with the design software ci, P, S(d ), calc and TDSP, S(-
d ), calc by a factor f SSPi for individual components or fSTDS for the TDS (Eqs. 5.210
and 5.210a). These discrepancies result from the differences between the nominal
and the minimum salt rejection rates of the manufacturer’s membrane elements.

ci,P,SðdÞ,eff ¼ ci,P,SðdÞ,calc  f SSPi : ð5:210Þ

ci, P, S(d ), eff ¼ effective concentration of component i in the system product at


design conditions [mg/l]
ci, P, S(d ), calc ¼ calculated concentration of component i in the system product at
design conditions [mg/l]
f SSPi ¼ salt passage safety factor of component i []

TDSP,SðdÞ,eff ¼ TDSP,SðdÞ,calc  f STDS : ð5:210aÞ

TDSP, S(d ), eff ¼ effective TDS in the system product at design conditions [mg/l]
TDSP, S(d ), calc ¼ calculated TDS in the system product at design conditions
[mg/l]
fSTDS ¼ TDS salt passage safety factor []
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 605

The values of the two safety factors f SSPi and fSTDS are determined by the number
of membrane elements in the RO unit as noted above, but they are also influenced by
design conditions such as the selected average specific flux of the system, the type of
membrane, and the method of precleaning. The range of values of the safety factor
for specific components such as chloride, bromide, and boron, f SSPi , may be about
1.15–1.35 and that of the TDS safety factor fSTDS around 1.2–1.5. For verifying the
performance of industrial-scale seawater desalination plants, membrane
manufacturers specify in their guarantee conditions that a value of 1.2 will not be
exceeded for the f SSPi of chloride and bromide, 1.2–1.3 for boron, and 1.2 for the
fSTDS of the TDS.
If these safety margins of the membrane manufacturers should also be reflected in
the design of the seawater desalination plant, this can be considered when designing
the second pass. Then the safety factors f SSPi and fSTDS shall be taken into account
when calculating the capacity factor fC, RO2 for post-desalination in accordance with
Sect. 4.2.1.4.1, Eqs. (4.41) and (4.43). Both these safety factors shall be taken into
account when determining the feed concentration to the second pass, the composi-
tion of the bypass, and the post-desalination design. The second pass with specifica-
tion of maximum concentrations of the chloride, bromide, or boron components in
the SWRO plant’s final product and with conventional configuration of the main
desalination tract is calculated with Eq. (5.211). If a limit value is specified for the
salt concentration (TDS) in the plant’s product, fC, RO2 is calculated with
Eq. (5.211a).

ci,PðdÞ,RO1  ci,P,M
f C,RO2 ¼
ci,PðdÞ,RO1  ci,PðdÞ,RO2
ci,PðdÞ,RO1,calc  f SSPi ci,P,M
¼ : ð5:211Þ
ci,PðdÞ,RO1,calc  f SSPi  ci,PðdÞ,RO2,calc  f SSPi

fC,RO2 ¼ capacity factor RO2 (RO second pass) []


ci,P,M ¼ concentration of constituent i in final product [mg/l]
ci, P(d ), RO1 ¼ concentration of constituent i in the first pass product at design
conditions [mg/l]
ci,P(d ),RO2 ¼ concentration of constituent i in the second pass product at design
conditions [mg/l]
ci, P(d ), RO1, calc ¼ calculated concentration of constituent i in the first pass product
at design conditions [mg/l]
ci, P(d ), RO2, calc ¼ calculated concentration of constituent i in the second pass
product at design conditions [mg/l]
606 5 Reverse Osmosis Membrane System: Core Process of SWRO

TDSPðdÞ,RO1  TDSP,M
f C,RO2 ¼
TDSPðdÞ,RO1  TDSPðdÞ,RO2
TDSPðdÞ,RO1,calc  f STDS  TDSP,M
¼ : ð5:211aÞ
TDSPðdÞ,RO1,calc  f STDS  TDSPðdÞ,RO2,calc  f STDS

TDSP,M ¼ concentration of TDS in final product [mg/l]


TDSP(d ), RO1 ¼ concentration of TDS in the first pass product at design conditions
[mg/l]
TDSP(d ),RO2 ¼ concentration of TDS in the second pass product at design
conditions [mg/l]
TDSP(d ), RO1, calc ¼ calculated concentration of TDS in the first pass product at
design conditions [mg/l]
TDSP(d ),RO2,calc ¼ calculated concentration of TDS in the second pass product at
design conditions [mg/l]
If the first pass is configured in split-partial mode, the capacity factor fC, RO2 for
limiting components is calculated with Eq. (5.211b) and if a limit value for the TDS
is specified in the finished product, Eq. (5.211c) applies.

ci,BðdÞ,RO2  ci,P,M
f C,RO2 ¼
ci,BðdÞ,RO2  ci,PðdÞ,RO2
ci,B,RO2ðdÞ,calc  f SSPi  ci,P,M
¼ : ð5:211bÞ
ci,B,RO2ðdÞ,calc  f SSPi  ci,PðdÞ,RO2,calc  f SSPi

TDSBðdÞ,RO2  TDSP,M
f C,RO2 ¼
TDSBðdÞ,RO2  TDSPðdÞ,RO2
TDSB,RO2ðdÞ,calc  f STDS  TDSP,M
¼ : ð5:211cÞ
TDSB,RO2ðdÞ,calc  f STDS  TDSPðdÞ,RO2,calc  f STDS

ci, B, RO2(d ) ¼ concentration of constituent i in bypass, second pass [mg/l]


ci, B, RO2(d ), calc ¼ calculated split partial mode concentration of constituent i in
bypass, second pass, at design conditions [mg/l]
TDSB, RO2(d ), calc ¼ calculated split partial mode TDS in bypass, second pass, at
design conditions [mg/l]
In each of these cases, the capacity factor fC,RO2 of the second pass increases,
i.e. the capacity of the post-desalination stage is elevated. This is due both to the
increased concentration of the limiting components due to the safety factors and the
increase in TDS in the feed to the second pass as well as in the bypass line of the
post-desalination stage.
When selecting the values for the safety factors f SSPi and fSTDS that are
incorporated into the equations for calculating the capacity factor, it must be taken
into account that these safety factors for the membrane system refer only to “clean”
membranes. The increases in component concentrations and salt content in the first
and second stage products due to fouling and scaling are not taken into account.
However, the second pass also has the task of compensating for such influences in
5.4 RO Membrane Process Design with Membrane Manufacturers’ Design Software 607

order to guarantee a largely constant quality of the SWRO system’s finished product.
It would therefore certainly be expedient to take higher values for the safety factors
in this case than what from the manufacturer’s viewpoint would be required for
securing membrane performance and thus to dimension the post-desalination stage
for a correspondingly higher capacity, possibly even for 100% of the product flow of
the RO system. Thus, both safety factors f SSPi and fSTDS should also include
proportionate allowances for the increase in component and TDS concentrations in
the product of the main desalination stage due to the CEOP effect as set out in Sects.
5.3.2.2.2 and 5.3.3.3, Eqs. (5.161) and (5.161a).
If a TDS target value TDSP, SWRO(d ), eff has been specified for the salinity of the
SWRO plant’s final product, it has to be borne in mind that this will be influenced by
its post-treatment for purposes of potabilization. According to Eqs. (5.212) and
(5.212a), allowance should be made for the resulting increase in salinity as early
as at the RO system’s design phase, which means that the main desalination stage
and the second pass must be designed so that the salt content of the product TDSP, S(-
d ), eff of the RO systems corresponds to the target TDS value TDSP, SWRO(d ), eff
minus the value ΔTDSP, posttr of the TDS increase in the post-treatment stage.

TDSP,SWROðdÞ,eff ¼ TDSP,SðdÞ,eff þ ΔTDSP,posttr : ð5:212Þ

TDSP,SðdÞ,eff ¼ TDSP,SWROðdÞ,eff  ΔTDSP,posttr : ð5:212aÞ

ΔTDSP, posttr ¼ increase of TDS during post-treatment of product [mg/l]

5.4.3.2 Product Capacity


Like for the standard salt rejection, membrane elements likewise have a certain
bandwidth in their standard product flow. The elements’ product flux can vary by
10–15% of the normalized value as stated in the specification. Here too, as
described above for salt rejection, product flow tends towards the normalized
value as laid down in the specification as the number of membrane elements in the
RO system increases.
To avert a reduction in the RO system’s performance, there are two options:

• raising the inlet pressure pF,S to the membrane system as calculated with the
manufacturer’s design program by a value of ΔpF, S, safety. This value is calcu-
lated, with inclusion of a corresponding safety factor fS,p,Pf, from the net driving
pressure of the NDPS system using Eq. (5.213).

NDPS  f S,p,Pf
ΔpF,S,safety ¼ : ð5:213Þ
100

ΔpF, S, safety ¼ RO system feed pressure increase for product flow loss compen-
sation [bar]
NDPS ¼ net driving pressure of RO system [bar]
608 5 Reverse Osmosis Membrane System: Core Process of SWRO

fS, p, Pf ¼ product flow safety factor for product flow loss, pressure compensation
[%]
¼ ~10–15%

• reserving space for additional membrane modules within the RO system array as
a precaution should its specified product flow not be achieved. The number of
backup membrane elements required for this is determined by applying the safety
factor fS,M,Pf according to Eq. (5.214) with the resulting number of membrane
modules calculated using Eq. (5.214a).

N ES  f S:M,Pf
ΔN ES,safety ¼ : ð5:214Þ
100
ΔN ES,safety
ΔN M S,safety ¼ : ð5:214aÞ
N E,M

N ES ¼ number of elements in RO array/system


ΔN ES,safety ¼ number of additional reserve elements []
fS. M, Pf ¼ product flow safety factor for product flow loss, membrane compensa-
tion [%]
¼ ~10–20%
NE, M ¼ number of membrane elements per module []
ΔN M S,safety ¼ number of additional reserve membrane modules []
In practice, the design of reverse osmosis systems usually favours the option of
allowing additional space for membrane modules in the array. In a multi-stage array,
the unoccupied back-up module slots are then distributed throughout the stages
depending on the number of modules in each stage.

5.5 RO Process Units, Components and Overall System


Configuration

5.5.1 RO Process Units

The principal process components of the reverse osmosis system of an RO seawater


desalination plant are shown in Fig. 5.63.
They consist of:

• the main RO desalination stage (first pass) with:


– the high-pressure (HP) pump group with energy recovery system
– the safety cartridge filter station
– dosing points for dosing chemicals into the feed of the first pass with their
preparation stations and dosing pump units
5.5 RO Process Units, Components and Overall System Configuration 609

– the first pass main desalination membrane system


– intermediate permeate storage tanks for the main desalination stage
• the RO post-desalination stage (second pass) with:
– the second pass feed pumps
– dosing points for dosing chemicals into the feed line of the second pass
– the second pass post-desalination membrane system
– concentrate recirculation from the second to the first pass
– a bypass line for the second pass if this is not dimensioned for the full RO
product flow (not shown in Fig. 5.63)
• equipment for chemical cleaning with rinsing and flushing of the membranes in
the first and second passes with:
– the preparation tank for the cleaning chemicals
– equipment for setting the optimum membrane cleaning temperature
– pumping station with the rinsing and flushing pumps
– a cartridge filter station for cleaning the rinsing solutions
• online instrumentation for recording the plant operating data and the RO control
and monitoring system

1st pass feed 2nd pass feed


dosing points dosing points

Ac As C Ss Ss Sh As

1st Pass
2nd Pass feed
Supply Cartridge High pressure Intermediate
pumps
pumps safety filters feed pumps permeate storage
Energy
Concentrate
recovery

2nd Pass
2nd pass Product
concentrate

Cleaning system
1st and 2nd pass

Dosing point code:


Ac = Acid
As = Antiscalant
C = Chlorine ( in case of CA membranes in 1st pass)
Ss = Sodiumbisulphite ( in 2nd pass feed in case of CA membranes in 1st pass)
Sh = Sodium hydroxide

Fig. 5.63 RO system configuration and components


610 5 Reverse Osmosis Membrane System: Core Process of SWRO

The HP pump group normally consists of two pumps, these being the actual HP
pump and a booster pump, which either provides the necessary inlet pressure for HP
pump operation or else serves to take over part of the overall pressure increase in the
RO feed line (see Sect. 5.5.2.2.2).
Depending on the type of energy recovery system installed, this is either
connected to the HP pump, connected to the entire first pass feed, or it transfers
the energy from the first pass concentrate directly to an RO feed section to boost the
pressure. In the latter case, the first pass feed flow is split into a part that goes to the
HP pump and another part that goes to the energy recovery system, which then feeds
it into the RO unit (see Sect. 5.5.2.2.3).
The dosing points for antiscalant, acid, chlorine, and bisulphite for dechlorination
in the feed to the main desalination stage (first pass) and in the feed to the second
pass of antiscalant and sodium hydroxide as well as of bisulphite for dechlorination
if cellulose acetate membranes are used in the first pass and polyamide membranes
for post-desalination are labelled accordingly in Fig. 5.63 (see Sect. 5.5.2.4).
The RO system layout of a large-scale seawater desalination plant with its
modular membrane separation units, the associated HP pumps, and energy recovery
systems together with the connecting piping between the components is shown in
Fig. 5.64.
In this case, all systems are housed in a reverse osmosis building. In other plants,
in part the RO membrane systems are separated from the very noisy HP pump
components and energy recovery devices.

Fig. 5.64 RO system with RO membrane module racks, high-pressure pumps, and energy
recovery systems (courtesy: Suez Water Technologies & Solutions)
5.5 RO Process Units, Components and Overall System Configuration 611

5.5.2 RO Process Components

5.5.2.1 RO Module Assembling Facilities


The membrane modules are arranged in racks of steel girders with protective paint or
other coating to prevent their corrosion should they come into contact with seawater
or concentrate. Figure 5.64a shows the arrangement of 800 modules in such a rack
from a side view. Also shown are the free slots for installing additional membrane
modules and the arrangement of two spare pressure vessels.
Figure 5.65b shows the front view of a module rack of a large RO seawater
desalination plant from the product extraction side with the permeate headers at the top.
The permeate headers are fed via a collection system consisting of supply lines
from each module to collection lines arranged between the modules. Permeate can be
removed from both ends of the membrane modules, this being a requirement for the
construction of the main desalination stage in the split-partial design (see
Sect. 5.2.3.2).
Each module is also connected to manifolds for feed admission to the membrane
modules and for discharge of the generated concentrate. These manifolds are in turn
connected to feed headers and concentrate headers.
The configuration of the headers and manifolds depends on the design of the
membrane modules. A distinction is made between end-port, side-port, and multi-
port designs.

Fig. 5.65 (a) RO module


arrangement for a high-
capacity RO system (source:
author). (b) RO module
arrangement in a high-
capacity RO system (Source:
author)
612 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.66 (a) RO module end-port connection (source: author). (b) RO module side-port connec-
tion (source: author). (c) RO module multi-port connection (Source: author)

For the end-port version, the feed to the membrane module and the discharge of
the concentrate are via nozzles in the modules’ end caps. In addition to the feed and
concentrate nozzles, the end caps also accommodate the permeate discharge nozzles.
In this design, feed admission and concentrate discharge are via manifolds arranged
in front of or between the modules (Fig. 5.66a). Manifolds for feed admission and
5.5 RO Process Units, Components and Overall System Configuration 613

concentrate discharge are also required if side-port modules are used. For this design,
feed to the modules and discharge of concentrate are via lateral nozzles in the module
jacket (Fig. 5.66b). The modules’ end caps contain only the two permeate discharge
nozzles. For connecting the multi-port membrane modules, these likewise have nozzles
for feed admission and concentrate discharge in the module jacket. These modules,
though, are connected directly to the headers on their feed and concentrate sides, so
there is no need for intermediate manifolds (Fig. 5.66c). In this case, feed and concen-
trate are distributed via the interconnected end pieces of the membrane modules.
The side-port and multi-port designs find application predominantly in large
SWRO plants, while the end-port design is preferred for small and medium-sized
installations. Depending on the module design variant, the membrane pressure
vessels and their distribution systems exhibit differing pressure losses. The
corresponding values for differential pressure must therefore be determined for the
range of operating conditions of the RO system by consulting the pressure vessel
manufacturers’ documentation and on this basis, while also factoring in design and
economic aspects, an appropriate membrane module model must be selected for the
design of the RO system.
Deviating from the hitherto common horizontal configuration of membrane
modules, the 1600 membrane elements used in a large SWRO plant were installed in
vertically arranged membrane pressure vessels (Sorek SWRO Plant, Israel) (Fig. 5.67).
The connections of the membrane modules as well as of the manifolds and
headers for feed admission and concentrate discharge in the HP rack section are,
depending on their diameter, of seamless or welded stainless steel pipes. High alloy
austenitic-ferritic (Super-Duplex) stainless steels are used as pipeline material (see
[13] Chap. 6 in Volume 2). The connecting lines are attached to the membrane
modules via grooved pipe couplings (Victaulic couplings). For larger diameters,
stainless steel flange connections are used.

Fig. 5.67 Vertical arrangement of 16" RO modules (Courtesy: IDE Technologies)


614 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.68 RO Module racks


with protection grating
(source: author)

Material specification, design, manufacture, and testing of the high-pressure


pipelines are based on national, international, or operator-specific standards.27 In
membrane desalination, it is mainly the ASME and ASTM standards and
specifications that are used as a basis for the calculation and construction of high-
pressure pipes and their pressure testing.
To protect the operating personnel, the RO racks are equipped with protection
grating on the front and at rear over the normal height of operational range
(Fig. 5.68).

27
ASME B31.3 Process Piping; ASME B36.10M Welded and Seamless Wrought Steel Pipe;
ASME B36.19M Stainless Steel Pipe; ASME B16.5 Pipe Flanges and Flanged Fittings; ASME
B16.9 Factory Made Wrought Butt-welding Fittings; ASTM A312/A312M Standard Specification
for Seamless and Welded Austenitic Stainless Steel Pipes.
EN 13480 European metallic industrial piping code; EN ISO 1127 Stainless steel tubes.
5.5 RO Process Units, Components and Overall System Configuration 615

5.5.2.2 High Pressure Feed Supply and Energy Recovery Systems

5.5.2.2.1 RO High Pressure Feed System: Pressure and Flow Design


Feed Pumps Delivery Pressure
In order to maintain a uniform product flow of the RO system even under fluctuating
operating conditions, the feed pump installation of the desalination plant has to be
dimensioned such that it is possible to reliably compensate for such influences by
adjusting the pressure when feed flow is admitted to the membrane system. Thus, for
the design of the HP pumps and their supply pump systems, first the bandwidths of
the parameters influencing the performance of the RO system must be determined.
These parameters are:

• the ranges of fluctuation of seawater temperature and salinity


• the degree of fouling to be expected and the resulting increase in the differential
pressure of the RO membrane elements, i.e. the maximum differential pressure in
the membrane modules up to which the system is operated before the membranes
are cleaned
• the design specification for average membrane lifetime (AMLT), i.e. the rate of
membrane replacement and the resulting increase in the operating pressure of the
membrane system from commissioning through to attainment of AMLT
equilibrium
• changes of the differential pressure in the pressure and suction lines of the feed
pumps due to filter systems installed there, like cartridge safety filters or mem-
brane filtration systems of the pretreatment system
• influences of the energy recovery system on the operating pressure required for
the membranes
• safety factors specified by the membrane manufacturer, the pump supplier, or the
designer to ensure safe operation of the HP pump systems.

With knowledge of these parameters, best-case and worst-case operating points


are defined, which then fix the bandwidth for the pressure rating of the HP feed
pumps.
The worst case operating point for the pressure rating is the feed pressure
required for the membranes at maximum seawater salinity, lowest seawater temper-
ature, maximum fouling of the membranes and the resulting pressure loss of the
membrane modules, operation at the specified average membrane lifetime, and the
highest pressure loss of filter systems integrated into the feed pumping systems.
The best case operating point is defined by minimum seawater salinity, highest
seawater temperature, clean membranes, the operating pressure for the RO
membranes during or after commissioning of the plant, and lowest differential
pressure of integrated filter systems.
The worst case operating point determines the highest pressure that the feed
pumps must deliver and the best case operating point the lowest pressure at which
the system has to be operated. Both operating points are then assigned appropriate
safety margins for the design of the HP and supply pumps.
616 5 Reverse Osmosis Membrane System: Core Process of SWRO

If the RO plant is to be operated at different recovery rates, additional parameters


to be included into design are the value of the highest product recovery rate at the
worst-case operating point, and the figure for the lowest recovery rate at the best-case
operating point.
The operating pressure pP, T(d ) for which the feed pumps are to be rated at the
respective operating point is calculated with Eq. (5.215). This is made up of the feed
pressure to the RO membrane system pF, T, RO(d ) for the design condition at the
operating point, the pressure loss Δppiping(d ) due to the pipes and fittings at the RO
infeed and in the plant itself, and, if an isobaric energy recovery system is installed,
the increase in the feed pressure to the membranes ΔpF, RO(d ), mix due to admixture of
the part stream from the ERD with its higher salinity. Further, the pressure difference
Δpgeo resulting from the geodetic height between the feed pumps themselves and the
RO rack must also be taken into account.

pP,T ðdÞ ¼ pF,T,ROðdÞ þ ΔppipingðdÞ þ ΔpF,ROðdÞ,mix þ Δpgeo : ð5:215Þ

pP, T(d ) ¼ HP feed pump installation pressure at design conditions [bar]


pF, T, RO(d ) ¼ pressure of RO membrane system feed at design conditions [bar]
Δppiping(d ) ¼ pressure loss caused by piping and valves [bar]
ΔpF, RO(d ), mix ¼ pressure increase at design conditions due to mixing of salinities
in the RO feed if isobaric ERD systems are installed [bar]
Δpgeo ¼ pressure increase from geodetic height difference from the hydraulic
profile [bar]
The total feed pressure to the RO modules pF, T, RO(d ) at the conditions of the
respective operating point, i.e. temperature, salinity; membrane age, and the extent
of membrane fouling results from the total feed pressure pF, T, RO(d ), clean required for
the “clean” but aged membranes; membrane fouling that has occurred during the
deposition time ΔτD and the therefore necessary increase in the feed pressure
Δp f ,T,ROðdÞ,ΔτD to compensate for the pressure drop in the flow channels; and the
CEOP effect at the walls of the membrane elements as well as the pressure loss ΔpPV
(d ) of the pressure vessels in which they are installed (Eqs. 5.215a and 5.215b).

pF,T,ROðdÞ ¼ pF,T,ROðdÞ,ΔτD þ ΔpPV ðdÞ : ð5:215aÞ

pF,T,ROðdÞ,ΔτD ¼ pF,T,ROðdÞ,clean þ Δp f ,T,ROðdÞ,ΔτD : ð5:215bÞ

pF,T,ROðdÞ,ΔτD ¼ necessary feed pressure to fouled membrane system at design


conditions [bar]
ΔpPV(d ) ¼ pressure loss caused by pressure vessels [bar]
pF, T, RO(d ), clean ¼ necessary feed pressure to the clean membrane system at
design conditions ¼ membrane feed pressure output by membrane manufacturers’
design software [bar]
Δp f ,T,ROðdÞ,ΔτD ¼ total feed pressure increase to membrane system at design
conditions due to fouling [bar]
5.5 RO Process Units, Components and Overall System Configuration 617

The total feed pressure required for the clean membranes pF, T, RO(d ), clean is made
up of the pressure component pF, RO(d ), clean needed to generate the required product
flux through the membranes together with the component ΔpFCðdÞclean to overcome
the pressure loss in the flow channels of the membrane elements (Eq. 5.215c). Both
pressure parameters are provided by the calculation results of the membrane
manufacturers’ design software.

pF,T,ROðdÞ,clean ¼ pF,ROðdÞ,clean þ ΔpFCðdÞclean : ð5:215cÞ

pF, RO(d ), clean ¼ necessary feed pressure to clean membrane system [bar] ¼ output
of membrane manufacturers’ design software [bar]
ΔpFCðdÞclean ¼ pressure loss of clean flow channels ¼ output of membrane
manufacturers’ design software [bar]
The increasing flow resistance Δp f ,T,ROðdÞ,ΔτD in the flow channels due to
membrane fouling during the deposition period ΔτD results from the higher flow
resistance arising from coating of the element spacers and the membrane surfaces as
well as the rise in osmotic pressure due to fouling, the so-called cake-enhanced
osmotic pressure CEOP (Eqs. 5.215d and 5.215e) (see Sects. 5.3.2.2.2 and
5.3.3.2.2). The increase in feed pressure Δp f ,T,ROðdÞ,ΔτD needed to compensate for
the fouling-induced product flow reduction Δπ CEOP,ΔτD and pressure loss likewise
due to fouling in the flow channels ΔpFC,fouledðdÞ,ΔτD is calculated with Eq. (5.215d).

Δp f ,T,ROðdÞ,ΔτD ¼ ΔpFC,fouledðdÞ,ΔτD þ Δπ CEOP,ΔτD : ð5:215dÞ


 
Δπ CEOP,ΔτD ¼ π F  CFðdÞ  βclean  F CEOP,ΔτD  1 : ð5:215eÞ

Δp f ,T,ROðdÞ,ΔτD ¼ total feed pressure increase to membrane system at design


conditions due to fouling [bar]
ΔpF,C,fouledðdÞ,ΔτD ¼ pressure loss of feed to concentrate of system flow channels
due to fouling at design conditions (d) over the deposition time ΔτD [bar]
Δπ CEOP,ΔτD ¼ CEOP induced increase of osmotic pressure [bar]
F CEOP,ΔτD ¼ CEOP induced average fouling factor over deposition time ΔτD []
In practice, however, determining the CEOP component of the fouling-induced
flow resistance of the membranes is laborious. If, though, the necessary feed pressure
to the membrane system after a defined operating time pF,T,ROðdÞ,ΔτD is set in relation
to the feed pressure pF, T, RO(d ), clean required to achieve the required product flow
with coated membranes so as to attain the same product flow as for “clean”
membranes; an empirical fouling factor F f ,T,M ðdÞ,ΔτD is obtained which takes into
consideration the pressure influences of both the coating and the CEOP (Eqs. 5.215f
and 5.215g). Both of these pressure parameters can be measured at various operating
points on plants while they are running and the fouling factor Ff,T,M,ΔτD can then be
calculated from this using Eq. (5.215f).
618 5 Reverse Osmosis Membrane System: Core Process of SWRO

pF,T,ROðdÞ,ΔτD
F f ,T,M ðdÞ,ΔτD ¼ : ð5:215fÞ
pF,T,ROðdÞ,clean

pF,T,ROðdÞ,ΔτD ¼ F f ,T,M ðd Þ,ΔτD  pF,T,ROðdÞ,clean : ð5:215gÞ

F f ,T,M,ΔτD ¼ factor of total reversible fouling over deposition time ΔτD []
From Eqs. (5.215g) and (5.215b), a relationship can be derived according to
Eq. (5.215h) with which the pressure loss of the coated membranes due to all fouling
components can be determined from the feed pressure of the “clean” membrane
system by applying the fouling factor F f ,T,M,ΔτD .
 
Δp f ,T,ROðd Þ,ΔτD ¼ pF,T,ROðdÞ,clean  F f ,T,M ðd Þ,ΔτD 1 : ð5:215hÞ

The pressure which the feed pump installation must provide to the RO system at
the operating point is then calculated according to Eq. (5.215i).

pP,T ðdÞ ¼ pF,T,ROðdÞ,clean þ Δp f ,T,ROðd Þ,ΔτD þ ΔpPVðdÞ þ ΔppipingðdÞ


þ ΔpF,ROðdÞ,mix þ Δpgeo : ð5:215iÞ

For design in practice of RO systems, dimensioning to take account of fouling


and calculation of the pump delivery pressure usually uses the increase in pressure
loss at the membrane modules ΔpF,CðdÞ,ΔτD , as measured for monitoring fouling at
the modules of operating plants. This encompasses the pressure loss due to fouling of
the flow channels ΔpF,C,fouledðdÞ,ΔτD as well as that of the clean membrane elements
ΔpFCðdÞclean (Eq. 5.215j).

ΔpF,CðdÞ,ΔτD ¼ ΔpFC,fouledðdÞ,ΔτD þ ΔpFCðdÞclean : ð5:215jÞ

ΔpF,CðdÞ,ΔτD ¼ total pressure loss (clean and fouled) in flow channels [bar]
According to the membrane manufacturers’ specifications, under operating
conditions this measurement should not increase to more than 15–25% of the
normalized pressure loss of a “clean” membrane module ΔpFCðdÞclean as calculated
by the membrane manufacturer’s software for a specific operating point
(Eq. 5.215k). However, it should be borne in mind that the value of ΔpF,CðdÞ,ΔτD
only relates to the pressure loss due to coating of the membranes’ flow channels, but
not the reduction in flow through the membranes as a result of their coating and the
influence of CEOP.

ΔpF,CðdÞ,ΔτD, max ¼ 1:15  1:25  ΔpFCðdÞclean : ð5:215kÞ

ΔpF,CðdÞ,ΔτD, max ¼ maximum differential pressure increase of membrane system at


design conditions due to fouling [bar]
5.5 RO Process Units, Components and Overall System Configuration 619

Using the value selected for ΔpF,CðdÞ,ΔτD, max , the design feed pump pressure may
then be calculated with Eq. (5.215l). pF, RO(d ), clean is calculated with Eq. (5.215c).

pP,T ðdÞ ¼ pF,ROðdÞ,clean þ ΔpF,CðdÞ,ΔτD, max þ ΔpPVðdÞ þ ΔppipingðdÞ


þ ΔpF,ROðdÞ,mix þ Δpgeo : ð5:215lÞ

If the feed pump installation consists of two series-connected pumps, namely a


supply or booster pump and the HP pump, generation of the feed pressure to the RO
unit is split proportionally between the two pumps according to Eq. (5.216).

pP,T ðdÞ ¼ pp,SPðdÞ þ pp,HPðdÞ : ð5:216Þ

pp, SP(d ) ¼ operating pressure of supply/booster pump at design conditions [bar]


pp, HP(d ) ¼ operating pressure of HP pump at design conditions [bar]
The required operating pressure pp, SP(d ) of the supply/booster pump results from
its inlet pressure, i.e. its static inlet head pSW, in, the desired inlet pressure to the HP
pump pF, HP, and pressure losses ΔpF, SP in the delivery line of the pump. If a safety
cartridge filter is arranged between the supply pump and the HP pump, the maximum
differential pressure of this filter plus the additional pressure loss in the pipes and
fittings of the pressure line of the booster pump (Eq. 5.216a) must be used for the
design to compensate for these pressure losses.

pp,SPðdÞ ¼ pF,HP þ ΔpF,SP  pSW,in : ð5:216aÞ

pSW, in ¼ suction line pressure of booster pump [bar]


ΔpF, SP ¼ pressure loss in feed piping of booster pump [bar]
pF, HP ¼ pressure to HP pump [bar]
The required operating pressure of the HP pump pp, HP is then calculated
according to Eq. (5.216b) from the sum of:

• the feed pressure pF, T, RO(d ) to the membrane system


• the pressure losses Δppiping(d ) in its piping network
• the geodetic height Δpgeo to be overcome by the feed pump installation
• if an isobaric energy recovery system is installed, the additional pressure ΔpF, RO
(d ), mix that results from the higher salinity in the feed to the RO unit

minus the intake pressure pF, HP to the HP pump as provided by its supply pump.

pp,HPðdÞ ¼ pF,T,ROðdÞ þ ΔppipingðdÞ þ ΔpF,ROðdÞ,mix þ Δpgeo  pF,HP : ð5:216bÞ

Flow of Feed Pumps


The range of feed flow FF,RO to the RO system is determined on the basis of its gross
design product capacity CGD and the values of the recovery rate YRO at which the
desalination system is operated. The gross design product capacity CGD is calculated
620 5 Reverse Osmosis Membrane System: Core Process of SWRO

from the net output product capacity CNO, i.e. the product capacity that the plant
delivers to the consumer, the plant’s internal consumption CP,internal, and its design
availability APD according to Eqs. (5.217) and (5.217a) (see Sect. 4.2.1.2.1).

C ND¼CNO 100 : ð5:217Þ


APD

C GD ¼ C ND þ CP,internal : ð5:217aÞ

CND ¼ net design product capacity [m3/day]


CNO ¼ net output product capacity [m3/day]
APD ¼ design availability [%]
CGD ¼ gross design product capacity [m3/day]
CP,internal ¼ internal product water consumption of plant or system [m3/day]
The required pumping capacity of the RO feed pump installation of RO1 depends
firstly on the selected type of energy recovery system and secondly on its pump-side
configuration, i.e. how many supply pumps and HP pumps have to be installed to
provide the feed flow to the system.
For energy recovery systems that are directly coupled to the HP pump or
integrated into the HP line of RO1, i.e. for Pelton turbines and turbochargers, the
feed flow FF,RO to the RO unit is the parameter from which the pump flows FP,RO1 of
the feed pump installation are derived.
From the capacity parameters calculated using Eqs. (5.217) and (5.217a), the
recovery rate of the system YRO is used to calculate the feed flow FF,RO to the RO
system, as shown by Eq. (5.218).

CGD C þ C P,internal 100  ðCNO þ C P,internal Þ


F F,RO ¼ ¼ ND ¼ : ð5:218Þ
24  Y RO 24  Y RO APD  24  Y RO
FF,RO ¼ feed flow to RO system [m3/h]
YRO ¼ product recovery factor of RO system
¼ YRO1 for one-pass system
¼ YRO1,RO2 for two-pass system
If the number N PF of supply pumps and HP pumps in the feed pump group is
equal (F P,RO1 ¼ F Psupply ,RO1 ¼ F PHP ,RO1 ), their pump delivery is calculated with
Eq. (5.218a).

F F,RO
F P,RO1 ¼ F Psupply ,RO1 ¼ : ð5:218aÞ
N PF

FP,RO1 ¼ pump flow to RO1 [m3/h]


N PF ¼ number of feed pump groups []
The relationship F P,RO1 ¼ F Psupply ,RO1 ¼ F PHP ,RO1, however, only applies to a pure
train configuration of RO1 if one supply pump is directly assigned to each HP pump.
In this case, the delivery of the pumps can also be calculated from the RO unit
product flow and its product recovery YRO in accordance with Eq. (5.218b). With
5.5 RO Process Units, Components and Overall System Configuration 621

this equation, it is also possible to determine what product flow FPr, RO1 can be
attained with one or more pumps in the feed pump installation and their respective
pump flows for a given product recovery rate YRO (Eq. 5.218c).

F Pr,RO
F P,RO1 ¼ F Psupply ,RO1 ¼ : ð5:218bÞ
N PF  Y RO
F P,RO1
F Pr,RO ¼ : ð5:218cÞ
N PF  Y RO

FPr, RO ¼ product flow of RO unit [m3/h]


If in the feed pump group the number of supply pumps differs from the number of
HP pumps, i.e. if the supply pumps and HP pumps are connected via manifolds, their
pump delivery is calculated with Eqs. (5.218d) and (5.218e).

F F,RO
F Psupply ,RO1 ¼ : ð5:218dÞ
N Psupply,RO1

F F,RO
F PHP ,RO1 ¼ : ð5:218eÞ
N PHP,RO1

F Psupply ,RO1 ¼ supply pump flow of RO1 [m3/h]


F PHP ,RO1 ¼ HP pump flow of RO1 [m3/h]
N Psupply,RO1 ¼ number of supply pumps []
N PHP,RO1 ¼ number of HP pumps []
If energy recovery takes place without direct coupling of the EDR with the HP
pump (isobaric systems), the pump delivery F PHP ,RO1 is then approximately the same
as the RO1 product flow FPr, RO1 (F PHP ,RO1 ffi F Pr,RO1 ) and the HP pump delivery is
then as shown by Eq. (5.218f).

F Pr,RO1 F F,RO  Y RO1


F PHP ,RO1 ffi ffi : ð5:218fÞ
N PHP,RO1 N PHP,RO1

For direct coupling, i.e. with Pelton turbine or turbocharger as ERD, F PHP ,RO1 is
calculated from the feed flow FF,RO to the RO unit using Eq. (5.218a) or from the
product flow of the RO unit FPr, RO with Eq. (5.218b).
An isobaric energy recovery system is supplied by a further part stream, whose
flow rate F FERD ,RO1 corresponds approximately to the feed flow to the RO system FF,
RO minus the product flow of RO1 (Eq. 5.218g).

F F ERD ,RO1 ffi F F,RO  F Pr,RO1 : ð5:218gÞ

F F ERD ,RO1 ¼ ERD supply flow [m3/h]


Such a system can be operated with one or two supply pump groups. When RO1
is operated with one group, its delivery F Fsupply ,RO1 corresponds to the feed flow to the
622 5 Reverse Osmosis Membrane System: Core Process of SWRO

RO system, with F Fsupply ,RO1 F F,RO and the individual pump delivery is given by
Eq. (5.218h).

F F,RO
F Psupplyt ,RO1 ffi : ð5:218hÞ
N Psupply ,RO1

N Psupply ,RO1 ¼ number of supply pumps for supply of HP—pump and ERD []
In the case of two supply pump groups, the one assigned to the HP pump group
has the same delivery as this. The pump delivery for the EDR supply pump group
F F ERD ,RO1 is given by Eq. (5.218g).
The deliveries of each supply pump are then calculated using Eqs. (5.218i) and
(5.218j).

F Pr,RO1
F P,supply,HP,RO1 ffi : ð5:218iÞ
N P,supply,HP,RO1

F PERD ,RO1
FP,supply,ERD,RO1 ffi : ð5:218jÞ
N P,supply,ERD,RO1

With isobaric energy recovery systems, the delivery rating of the HP pump must
also take into account the so-called lubricant or leakage flow. For this reason, the
supply flow of the HP pumps does not exactly correspond to the product flow of
RO1. If the isobaric ERD is operated with so-called “overflush”, the supply flow into
the ERD for this purpose, i.e. the delivery of the ERD supply pump, must still be
added to the feed flow to the RO system. Details of how to calculate isobaric energy
recovery systems can be found under Sect. 5.5.2.2.3.

5.5.2.2.2 RO High Pressure Feed Pumps: Type of Pumps, Their Configuration,


Efficiency, and Operation Control
Type of RO High-Pressure Pumps
In the feed pump installation of an RO seawater desalination system, both positive
displacement pumps and reciprocating pumps in the form of multiple-plunger pumps
as well as multi-stage, high-speed centrifugal pumps are used for the HP pump
range. Predominantly single-stage process centrifugal pumps are used for supplying
the HP stage of the feed pump installation.
Multiple-plunger pumps for the HP supply for reverse osmosis have deliveries
of between a few 100 l/h and up to 50–150 m3/h, depending on the manufacturer and
their design. In combination with an isobaric energy recovery system, a product flow
of 3600 m3/day can be achieved at this maximum delivery rate, and correspondingly
less with non-isobaric energy recovery to match the recovery rate of the RO
installation (Eq. 5.218f). This type of HP pump is therefore mainly used in smaller
desalination plants.
In the design shown in Fig. 5.69, the pumps are equipped with 3 or 4 plungers,
i.e. triplex and quadruplex pumps.
5.5 RO Process Units, Components and Overall System Configuration 623

Fig. 5.69 Triplex plunger pump (Courtesy: Cat Pumps)

Table 5.24 Plunger pumps—number of pistons/plungers of pump and related flow pulsation (Data
source: Danfoss)
Number of pistons/plungers 3 5 7 9
Degree of flow pulsation (%) 14 5 2.5 1.5

All reciprocating and plunger pumps generate a delivery that fluctuates in both
flow rate and pressure. The fluctuation amplitude depends on how many pistons
move during pump operation. Especially for pumps with fewer pistons or plungers,
the pump installation has to be fitted with a pulsation damper in its discharge line and
the pump’s suction and discharge nozzles have to be flexibly connected to their
intake and delivery lines so that the pump’s vibrations will not be transmitted to the
pipelines and other plant components.
The more pistons or plungers a pump has, the less is the pulsation amplitude in the
delivery flow (see Table 5.24).
Axial piston pumps (APPs), a type of reciprocating pump, contain up to nine
plungers that are arranged in a cylinder block parallel to the pump drive shaft so that
they rotate with it. They are in contact with the inclined surface of the fixed swash
plate which thus forces them to reciprocate while rotating. (Fig. 5.70). The stroke
length is fixed by the angle of inclination of the swash plate. A valve plate, which is
also fixed, controls the supply to the pistons in the cylinder block and the discharge
of the pressurized liquid. Thanks to the reduced flow pulsation with this type of
displacement pump, pulsation dampers are not required. Nevertheless, flexible
connectors to the inlet and delivery lines are still needed to prevent transmission
of vibrations.
Reciprocating/plunger pumps are usually operated in the RPM range of
1200–1800 min1.
624 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.70 Axial piston pump operating principle and pump construction (Courtesy: Danfoss)

For smaller, but also for medium-sized RO desalination systems, centrifugal


pumps are used whose construction is based on the design characteristics of well
pumps. This compact type of pump has a casing with a diameter that is small but it is
relatively long compared with conventional centrifugal pumps. This design is
achieved by using small-diameter impellers and a large number of stages that may
be as many as 20 for HP pumps and operation at RPM of 4000–6000 min1
(Fig. 5.71).
Depending on the manufacturer, this type of pump attains delivery rates in the HP
range of from 100 to 300 m3/h, corresponding to product flows of 2400–7200 m3/
day with isobaric energy recovery. These pumps can be used for both supply and HP
applications and, thanks to their very compact width albeit elongated design, are well
suited for application in prefabricated standardized RO systems.
In large-scale plants, primarily multi-stage centrifugal pumps with radial split ring
sections (Fig. 5.72) or axial split housings (Fig. 5.73) find application in the HP stage
of the feed pump installations.
Multi-stage HP centrifugal pumps in split-ring design achieve deliveries of up to
1500 m3/h, while centrifugal pumps with axially split housings achieve deliveries of
up to 3000 m3/h, depending on the feed pressure that the HP pump stage of the feed
5.5 RO Process Units, Components and Overall System Configuration 625

Fig. 5.71 High-pressure centrifugal pump-High speed multi-stage type (Courtesy: Grundfos)

Fig. 5.72 High-pressure centrifugal pump-Ring section multi-stage type (Courtesy: #KSB SE &
Co. KGaA)

pump installation has to provide for the RO system. Depending on the required flow
rate and pressure head, the pumps have from two to six stages and operate at RPM of
3000 min1. With these types of centrifugal pumps, with the maximum pump
capacities specified by the manufacturers for isobaric energy recovery, product
flows of around 36,000–72,000 m3/day are possible. If energy recovery is coupled
directly to the HP supply to the RO system via a Pelton turbine or turbocharger, at a
recovery rate of 45%, according to Eq. (5.218f) product flows of some 16,000 to
32,000 m3/day are possible.

Efficiency of RO Feed and Supply Pumps


The power demand of a pump unit PD,PX is calculated from its required delivery F PX
and its pressure head ΔH according to Eq. (5.219). Other parameters needed for the
calculation are the density of the pumped medium ρF, the standard gravitational
acceleration g, and the overall efficiency of the pump installation ηPX,T .
626 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.73 High-pressure centrifugal pump-Axial split multi-stage type (Courtesy: Sulzer—Two
stages, single suction, axial split casing pump model MSD-RO used as high pressure pump in
seawater reverse osmosis desalination applications)

F PX  ΔH  ρF  g
PD,PX ¼ : ð5:219Þ
3:6  103  ηPX,T

PD,PX ¼ power demand of pump X [kW]


F PX ¼ delivery of pump X [m3/h]
ΔH ¼ pressure head [m]
ρF ¼ density of pump feed [kg/l]
g ¼ standard gravitational acceleration ¼ 9.80665 [m/s2]
ηPX,T ¼ overall efficiency factor of pump set X []
This relationship applies to both positive displacement pumps and rotodynamic,
i.e. centrifugal, pumps. In piston/plunger pumps, the delivery is fixed by the piston
displacement, which is the piston/plunger diameter times the stroke length, as well as
the number of strokes per unit of time that depends on the speed at which the pump is
operated.
If the pump set’s power demand PD,PX is referred to the differential pressure Δp
which it generates in the system, Δp is calculated from the pressure head ΔH
according to Eq. (5.219a) and power demand PD,PX is then calculated with
Eq. (5.219b).
5.5 RO Process Units, Components and Overall System Configuration 627

ΔH  ρF  g
Δp ¼ : ð5:219aÞ
102
F PX  Δp
PD,PX ¼ : ð5:219bÞ
36  ηPX,T

Δp ¼ differential pressure [bar]


The overall efficiency of the pump set ηPX,T is made up of the individual
efficiencies of the pump ηP, motor ηM and, if the pump is equipped with variable
speed control, also the efficiency of the speed control drive ηVSD (Eq. 5.219c).

ηPX,T ¼ ηP  ηM  ηVSD : ð5:219cÞ

ηP ¼ efficiency factor of pump []


ηM ¼ efficiency factor of motor []
ηVSD ¼ efficiency factor of variable speed drive []
By substituting the algorithm Eq. (5.219c), that expresses the relationship of the
individual efficiencies to the overall pump efficiency, into Eq. (5.219b), the result is
Eq. (5.219d) for the power demand of the pump installation PD,PX , with the individ-
ual efficiencies ηP, ηM, and ηVSD of its components taken into account.

F PX  Δp
PD,PX ¼ : ð5:219dÞ
36  ηP  ηM  ηVSD

The feed pump installations exhibit the highest energy consumption of all
systems of an RO desalination plant. However, as Eq. (5.219d) shows, the energy
efficiency of the pump sets can be raised by optimizing the efficiencies of their
components. This applies in particular to the internal efficiency ηP of the
installation’s supply and HP pumps and thought must be given to their optimization
right from the design phase. This includes selection of the most suitable pump types
for the RO product flow as well as how pump delivery and pressure head of the feed
pump installation may be so controlled that the pumping units operate as closely as
possible to their best efficiency point (BEP) during plant operation.
When piston/plunger pumps are used to pump seawater, their pump efficiency is
in the 88–95% range, i.e. an efficiency factor ηP of 0.88–0.95. The pump efficiency
of larger pump sets, though, is generally independent of their delivery and pressure
head and also there is no significant difference in the efficiency bandwidths between
different types of reciprocating pumps.
This is not the case, though with centrifugal pumps. The design of the pumps, in
particular the type and design of the impellers and the nature of the flow through the
pump casing, significantly influence pump efficiency ηP. A centrifugal pump is
characterized by its specific speed nspec. This parameter defines the type and geome-
try of the impeller and means that pumps with the same value of nspec,
i.e. geometrically similar pumps, can be compared even if they have different
operating conditions, like delivery, pressure head, and speed.
628 5 Reverse Osmosis Membrane System: Core Process of SWRO

The specific speed is calculated from the delivery FP, BEP, the stage pressure head
HP, stage, BEP, and the speed nP, BEP of the centrifugal pump at the operating point for
which its efficiency is highest, i.e. the best efficiency point (BEP), according to
Eq. (5.220). For a multi-stage pump, HP, stage, BEP is calculated as the quotient of the
pump’s total pressure head HP, BEP at the BEP and its number of stages NP, stages
(Eq. 5.220a).
pffiffiffiffiffiffiffiffiffiffiffiffiffi
F P,BEP
nspec ¼ nP,BEP   34 : ð5:220Þ
H P,stage,BEP

H P,BEP
H P,stage,BEP ¼ : ð5:220aÞ
N P,stages

nspec ¼ specific speed [min1]


nP, BEP ¼ pump speed at operating point of best efficiency (BEP) [min1]
FP, BEP ¼ pump delivery at BEP [m3/s]
HP, stage, BEP ¼ head of one pump stage at BEP [m]
HP, BEP ¼ pump head at BEP [m]
NP, stages ¼ number of stages of multi-stage pump []
The multi-stage HP centrifugal pumps and single-stage process pumps of an RO
system are equipped with radial impellers. Depending on impeller design, specific
speeds nspec up to a maximum of 70 min1 can be attained with this type of pump.
The HP centrifugal pumps of an SWRO plant have specific speeds of approximately
30 min1. The single-stage supply pumps run in a specific speed range of
20–40 min1 and the booster pumps of the energy recovery system are in a range
of 40–60 min1. The feed pumps of the second pass of such a system have specific
speed values of around 20 min1.
The efficiency ηP of a centrifugal pump is significantly influenced by its specific
nspec speed and thus also, in accordance with Eq. (5.220), by its delivery FP, BEP. The
dependence of the theoretically possible efficiency of an HP four-stage vertically
split ring-section centrifugal pump on the specific speed and delivery rate is shown in
Fig. 5.74 [131].
Accordingly, for a centrifugal pump design with a certain specific speed, its pump
efficiency ηP increases disproportionately with delivery, particularly in the low to
medium delivery range. In the high delivery range it still increases, albeit to a lesser
degree. The internal pump efficiency attains an optimum in the range of specific
speeds of between 25 and 50 min1 depending on the delivery. However, the
theoretical values for pump efficiency of up to 90% shown in Fig. 5.74 are not
achieved in practice as there are other efficiency losses, in particular due to the
surface roughness of the impeller blades and pump casings and other construction-
related influences. Efficiency is also influenced by pump design, for example in the
HP range whether the pump is of the radial split ring section type or has an axially
split casing.
An algorithm for the dependence of the minimum efficiency on the centrifugal
pump models available on the market for water supply on their specific speed and
5.5 RO Process Units, Components and Overall System Configuration 629

Pump efficiency factor


ηP

Fig. 5.74 Pump internal efficiency factor as a function of specific speed

delivery is included in EU Regulation No. 547/2012 of 25 June 201228 (Eqs. 5.221,


5.221a, and 5.221b).
 
ηP,BEP, min requ ¼ 88:59  ln nspec þ 13:46  ln ðF P,BEP Þ  11:48
  2
 ln nspec  0:85  ½ ln ðF P,BEP Þ2  0:38
 ln ðF P,BEP Þ  C: ð5:221Þ

ηP,PL, min requ ¼ 0:947  ηP,BEP, min requ : ð5:221aÞ

ηP,OL, min requ ¼ 0:985  ηP,BEP, min requ : ð5:221bÞ

Validity: nspec ¼ 6–120 [min1]; FP, BEP ¼ 2–1000 [m3/h]; ηP, BEP, min requ ¼ 88%
This system of equations is based on the statistical evaluation of the perfor-
mance parameters of centrifugal pumps that were made available by their
manufacturers in 2007 in connection with a project to develop the aforementioned
EU regulation [132]. The factor C characterizes the differences in efficiency
between individual pump models and the technical level of their design, which is
characterized by the minimum efficiency index MEI. According to Table 5.25, the
factor C depends on the pump model together with its speed and MEI. The higher the
MEI, the better the quality of the pump’s technical design and thus its potential efficiency.

28
COMMISSION REGULATION (EU) No. 547/2012 of 25 June 2012 implementing Directive
2009/125/EC of the European Parliament and of the Council with regard to eco design requirements
for water pumps.
630

Table 5.25 C-values at minimum efficiency index MEI and pump speed for different types of pumps
Minimum Efficiency Index MEI
0.10 0.20 0.30 0.40 0.50 0.60 0.70
5

Type of pump Pump speed [min1] C-value


Singled stage end suction own bearing ESOB 1.450 132.58 130.68 129.35 128.07 126.97 126.0 124.85
2.900 135.60 133.43 131.61 130.27 129.18 128.12 127.06
Single stage end suction close coupled ESCC 1.450 132.74 131.20 129.77 128.46 127.38 126.57 125.46
2.900 135.93 133.82 132.23 130.77 129.86 128.80 127.75
Single stage end suction close coupled inline ESCCI 1.450 136.67 134.60 133.44 132.30 131.00 130.32 128.98
2.900 139.45 136.53 134.91 133.69 132.65 131.34 129.83
Vertical multi-stage MS-V 2.900 138.19 135.41 134.89 133.95 133.43 131.87 130.37
Submersible multi-stage MSS 2.900 134.31 132.43 130.94 128.79 127.27 125.22 123.84
ηP, BEP, min requ ¼ minimum efficiency at best efficiency point (BEP) [%]
ηP, PL, min requ ¼ minimum efficiency at part load (PL) point [%]
ηP, OL, min requ ¼ minimum efficiency at overload (OL) point [%]
FP, BEP ¼ pump delivery at BEP [m3/h]
C-value ¼ factor dependent on pump type, pump speed, and MEI
Reverse Osmosis Membrane System: Core Process of SWRO
5.5 RO Process Units, Components and Overall System Configuration 631

An MEI of 0.1 means that at the time the performance parameters were recorded in
2007, 10% of the pumps had not attained this C value. An MEI value of 0.4
corresponds to a proportion of 40% at the time the parameters were recorded.
According to [132], Eq. (5.221) for deriving the minimum efficiency at the BEP
together with the associated equations of the regulation relating to the operating
points at part load (Eq. 5.221a) and overload (Eq. 5.221b) apply for specific speeds
of 6–120 min1 with deliveries of 2–1000 m3/h and up to an efficiency of 88%.
EU Regulation No. 547/2012 stipulates that for the pump types listed in
Table 5.25 and depending on their performance parameters as prescribed in the
Regulation, the C-values stated for an MEI of 0.4 are to be used as the basis for
calculating the minimum efficiency ηP, BEP, min requ from 2015 onwards.
The HP pumps installed in RO seawater desalination plants are mostly not or only
to a limited extent covered by this Regulation with regard to design and performance
parameters and the other pump installations of large SWRO plants are only partially
covered. However, the algorithms of [132] and the EU Regulation make clear how
centrifugal pump efficiency depends on delivery and specific speed.
The Hydraulic Institute’s HI 20.3-2015 guideline29 can also be used to determine
the potential efficiencies of different types of centrifugal pumps. Thus, the influence
of surface roughness is also factored into the calculation, while the influence on
efficiency of the make-up of the medium to be pumped, the centrifugal pump model
together with its geometry and mechanical components are described in detail.
Figure 5.75 shows a comparison of the potential efficiencies of HP centrifugal
pumps in vertically split ring-section design versus axially split casings with differ-
ing numbers of stages [131]. It has been shown that with an optimum number of
stages and a delivery of 2000 m3/h, an efficiency of up to 87% can be achieved with
these types of pump.

Motor and Drive Efficiency


The power demand of a pump installation ηPX,T is influenced not only by the
efficiency of the pump ηP, but also by that of the drive motor ηM and other drive
components, for example a speed control system ηVSD (Eq. 5.219c).
Drive motors are assigned efficiency classes from IE1 to IE4 on the basis of
International Standard IEC60034-30 issued by the International Electrotechnical
Commission.30 Within the respective class, minimum motor efficiencies at rated
output power are defined, and these increase in step with motor power. The higher
the efficiency class of a motor, the greater is the minimum efficiency specified for it
(see Fig. 5.76).

29
HI 20.3-2015—Hydraulic Institute Guideline for Rotodynamic Pump Efficiency Prediction.
30
IEC 60034-30-1—Rotating electrical machines—Part 30-1: Efficiency classes of line operated
AC motors.
632 5 Reverse Osmosis Membrane System: Core Process of SWRO

Pump efficiency factor


ηP
0.88

0.87

0.86
3
2
5
0.85
4

4
0.84

0.83

3
5 Axial split pump with
0.82 X
X stages
2
Vertically split pump
X
with X stages
0.81
200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200

Pump flow rate [m3/h]

Fig. 5.75 Dependence of pump efficiency on pump flow rate of centrifugal pumps

Motor efficiency [%]


100

99
Motor efficiency classifications according to IEC 60034-30
98

97
IE4
96 IE3

95 IE2

94 IE1

93
IE1 = Standard efficiency
92
IE2 = High efficiency
91
IE3 = Premium efficiency
90
IE4 = Super premium efficiency
89

88

87
0 50 100 150 200 250 300 350
Motor power output [kW]

Fig. 5.76 Motor efficiency classes and dependence of efficiency on motor power output (for 50 Hz
4-pole motor)
5.5 RO Process Units, Components and Overall System Configuration 633

As Fig. 5.76 shows, in each of the efficiency classes the motors attain their
maximum efficiency at a rated output power of about 200 kW. Even with a further
increase in motor output, this efficiency remains the same and, depending on the
efficiency class, is in the range of 94–97%, i.e. a ηM value of between 0.94 and close
to 0.97.
According to EU Regulation No. 640/2009,31 as of 2017 all motors with a rated
output power of up to 375 kW should either achieve at least the efficiency level of
efficiency class IE3 or correspond to the efficiency level of class IE2 and be
equipped with speed control devices.
With speed control devices such as frequency converters, likewise efficiencies of
between 94% and 98% are attained. However, for both the motor and if a speed
control device is used, the minimum efficiencies as defined in the efficiency classes
are achieved only at the rated shaft power or at the rated drive speed. If motors run at
underload, there is a noticeable deterioration in efficiency. If frequency inverters are
used for speed control, this drop in efficiency is less pronounced, at least in the range
down to about 30% underload.

Net Positive Suction Head Requested (NPSHr) and Available (NPSHa)


Within an RO feed pump installation, at each pump set there must be sufficient head,
or pressure, on the intake side to prevent cavitation within the pump. The parameter
that characterizes the necessary inlet pressure conditions of a pump is the net positive
suction head requested, NPSHr.
The manufacturers of piston and plunger pumps specify the required NPSHr to
match the particular operating conditions of the pump in the form of the inlet
pressure that will then be required.
For a given plunger geometry, the NPSHr is primarily dependent on the speed at
which the pump is operated and thus on its delivery. If the necessary inlet pressure to
the piston pump cannot be provided by its head on the suction side, an additional
booster pump is needed for which purpose centrifugal pumps are normally used.
For centrifugal pumps, the NPSHr figure is one of its characteristic values and it is
plotted in a graph as a function of the pump’s delivery or, if the pump speed can be
varied, for different speeds. Pump manufacturers usually quote the NPSHr,3% value
for this purpose, this being the value of the NPSHr at which the pressure head of the
pump is reduced by 3%. The NPSHr,3% of a given pump depends on its delivery, the
associated pump speed, and the nature of the pumped medium.
Since the NPSHr,3% value of a centrifugal pump increases with operating time,
i.e. its suction performance deteriorates, an NPSHr,40,000 value is defined to take this
into account, which is calculated with Eq. (5.222) from the value of NPSHr, 3%. This
is an empirical value for the NPSHr, 3% that becomes established after a pump
operating time of 40,000 h on the basis of the manufacturer’s experience [133].

31
COMMISSION REGULATION (EC) No 640/2009 of 22 July 2009 implementing Directive
2005/32/EC of the European Parliament and of the Council with regard to eco design requirements
for electric motor.
634 5 Reverse Osmosis Membrane System: Core Process of SWRO

NPSHr,40,000h ¼ f NPSH,τoperation  NPSHr,3% : ð5:222Þ

NPSHr, 40,000h ¼ net positive suction head requested for 40,000 operation hours
[m]
NPSHr, 3% ¼ net positive suction head requested at 3% head loss [m]
f NPSH,τoperation ¼ NPSHr3% operation time increase factor ¼ 1.5–2.0
The feed conditions corresponding to the specific operating conditions of a
centrifugal pump in a pump installation are characterized by the value of the net
positive suction head available, NPSHa. This parameter is calculated with
Eq. (5.223). Depending on whether the pump is in suction head or suction lift
operation, the suction side geostatic height difference ΔHgeo is stated with a positive
or negative sign.

ps þ pa  pv,F
NPSHa ¼ 105   ΔH s ΔH geo,s : ð5:223Þ
ρF  g

+ΔHgeo, s ¼ suction head operation


ΔHgeo, s ¼ suction lift operation
NPSHa ¼ net positive suction head available [m]
ps ¼ pressure in suction line [bar]
pa ¼ atmospheric pressure [bar]
pv,F ¼ vapour pressure of feed at max. operating temperature [bar]
ρF ¼ density of feed [kg/m3]
g ¼ standard gravitational acceleration ¼ 9.80665 [m/s2]
ΔHs ¼ head loss in suction line [m]
ΔHgeo, s¼ height difference according to hydraulic profile between suction water
level and pump suction input level [m]
If seawater is the pumped medium, its vapour pressure pv,F is determined
according to Sects. 3.2.1 and 3.2.1.4, Eqs. (3.32), (3.33), (3.34), Fig. 3.12, and
Table 3.20, and its density ρF with Sects. 3.2.2 and 3.2.2.1, Eqs. (3.39), (3.40),
Fig. 3.15, and Table 3.23.
To keep cavitation in a pump to a minimum, the NPSHa must be greater than
NPSHr,3% (Eq. 5.223a).

NPSHa > NPSHr,3% : ð5:223aÞ

The ratio of NPSHa to NPSHr,3% depends on the pump characteristics and its
operating conditions. The higher the NPSHr,3% is at the operating point with the best
efficiency (BEP), the greater the safety factor f NPSHsafety should be [134] (Eq. 5.223b).

NPSHa  NPSHr,3%  f NPSHsafety : ð5:223bÞ

f NPSHsafety ¼ NPSHr,3% safety factor ¼ 1.2–1.5


5.5 RO Process Units, Components and Overall System Configuration 635

Information and calculation instructions for determining the NPSH safety factor
for different pump types and operating conditions are also contained in the Hydraulic
Institute’s ANSI/HI 9.6.1-2012 Guideline.32
For a centrifugal pump with speed control, the value specified for the safety factor
f NPSHsafety must also be maintained under the most unfavourable conditions for the
pump’s suction behaviour, these being maximum pressure loss ΔHs in the suction
line as well as the highest inlet temperature and thus also highest vapour pressure pv,
F of the pumped medium. This is especially true if cartridge safety filters or, as part of
the pretreatment stage, membrane filtration systems are installed in an RO feed pump
installation between the supply pump and the HP pump. In this case, the value of the
safety factor f NPSHsafety must be sufficient to allow even for the maximum pressure
drop across these units.
Derived from Eqs. (5.223) and (5.223b is an algorithm in accordance with
Eq. (5.223c) with which the required minimum inlet pressure ps for a centrifugal
pump can be calculated under its specific installation and operating conditions to
prevent or minimize cavitation.

ρF  g  
ps ¼  NPSHr,3%  f NPSHsafety þ ΔH s ΔH geo,s  pa þ pv,F : ð5:223cÞ
105

Pump System Configuration and Operation Control


The design of the RO feed pump installation must allow the product flow to be
maintained under the influence of temperature and salinity changes in the feed and
reduction of membrane permeability due to ageing and fouling as well as for varying
pressure conditions in the feed lines and membrane elements. These influences are
compensated for by adjusting the delivery head of the feed pumps. If the RO plant is
to operate with a variable recovery rate, its feed delivery rate must also be adjustable.
With piston/plunger pumps, the feed pressure adjusts itself independently of
pump delivery due to the flow resistance of the membrane system. In this case, the
pressure in the feed line to the RO unit is set by means of a concentrate control valve
installed in the plant’s concentrate-discharge line. In order to protect the system from
exceeding the permissible maximum operating pressure, for this type of pump a
safety valve and a safety pressure switch must be installed in the feed line to the RO
unit. Pump delivery is adjusted by controlling the speed.
With centrifugal pumps, the delivery rate and head are adjusted according to the
curve of the pump characteristic. During pumping operation, these two parameters
can be controlled by:

• throttling the feed by means of a throttle valve in the pump’s discharge line, or
• varying the pump speed.

32
ANSI/HI 9.6.1-2012—Rotodynamic Pumps Guideline for NPSH Margin.
636 5 Reverse Osmosis Membrane System: Core Process of SWRO

With throttle control, however, the energy consumption of a pump unit is


considerably higher than when its performance parameters are adjusted by speed
control. Depending on the pump design, it may even be necessary in the case of
throttling to recirculate part of the pump delivery back to the pump intake via a
bypass line. Hence, for seawater desalination plants, for both low product flows and
in large plants, today predominantly speed control equipment is installed at centrif-
ugal pumps to match RO operation to changing conditions.
When piston pumps are used for the high-pressure range, centrifugal pumps
primarily serve to generate an upstream pressure to prevent cavitation in the positive
displacement pump. The situation is different when centrifugal pumps are used in the
feed pump installation for both high-pressure applications and as supply/booster
pumps. In this case, the supply pump also contributes pressure generation as part of
the total feed pressure to the RO unit. It is then responsible for changing the
operating pressure and delivery to compensate for the varying influences on reverse
osmosis operation. For this purpose, the supply pump is equipped with infinitely
variable speed control and the HP pump is then operated mostly at its best operating
point with constant pressure and delivery. The operating pressure and concentrate
discharge of the RO unit are fine-tuned by control valves in the feed and concentrate-
discharge lines.
If, though, there are wide fluctuations in temperature or salinity in the feed to the
desalination plant, it is possible that the speed control of the supply pump alone can
no longer compensate for these influences. It may then become necessary to addi-
tionally equip the HP pump with speed control. In addition to modifying the
performance values of the feed pumps, product throttling, i.e. the generation of an
increased product-side pressure by means of a throttle valve in the product line, can
also contribute to adapting the system’s product flow to external influences (see
Sects. 5.2.3.2 and 5.2.3.2.2).
Varying the performance parameters of a centrifugal pump by controlling its
rotational speed can be modelled by applying the affinity law for similar centrifugal
pumps. Accordingly, the delivery FP of a centrifugal pump varies in direct propor-
tion to its speed (Eq. 5.224) and that of its delivery head HP with the square of the
speed nP (Eq. 5.224a).

F P,2 nP,2 n
¼ ; F ¼ F P,1  P,2 ð5:224Þ
F P,1 nP,1 P,2 nP,1

FP, 1 ¼ pump flow at operation point 1 [m3/h]


FP, 2 ¼ pump flow at operation point 2 [m3/h]
nP, 1 ¼ pump speed at operation point 1 [min1]
nP, 2 ¼ pump speed at operation point 2 [min1]
 2  2
H P,2 nP,2 nP,2
¼ ; H P,2 ¼ H P,1  : ð5:224aÞ
H P,1 nP,1 nP,1

HP, 1 ¼ pump head at operation point 1 [m]


5.5 RO Process Units, Components and Overall System Configuration 637

HP, 2 ¼ pump head at operation point 2 [m]


The pump’s power demand varies approximately with the cube of its speed
(Eq. 5.224b).
 3  3
PD,P,2 nP,2 nP,2
; PD,P,2 PD,P,1 : ð5:224bÞ
PD,P,1 nP,1 nP,1

PD, P, 1 ¼ power demand at pump shaft at operation point 1 [kW]


PD, P, 2 ¼ power demand at pump shaft at operation point 2 [kW]
Pump efficiency is likewise a function of the change in speed, as shown by
Eq. 5.224c.

  n
0:1
ηP,2 1  1  ηP,1  P,1 : ð5:224cÞ
nP,2

ηP, 1 ¼ pump efficiency factor at operation point 1 []


ηP, 2 ¼ pump efficiency factor at operation point 2 []
How the change in speed and the resulting change in efficiency affect the power
consumption of the pump is shown by Eq. (5.224d).
 
PD,P,2 ηP,1 n
3
ηP,1 n
3
 P,2 ; PD,P,2 PD,P,1   P,2 : ð5:224dÞ
PD,P,1 ηP,2 nP,1 ηP,2 nP,1

Also, the pump’s required NPSHRP is changed by controlling its speed according
to Eq. (5.224e) within its range of validity, as shown.
 x
NPSHRP,2 nP,2
ð5:224eÞ
NPSHRP,1 nP,1

x ¼ 2 for nP,2 ¼ 0.8–1.2  FP,opt at nP,1 and nspec < 106


NPSHRP, 1 ¼ net positive suction head requested at operation point 1 [m]
NPSHRP, 2 ¼ net positive suction head requested at operation point 2 [m]
The characteristics for delivery head, NPSHR, and power demand of a supply
pump operating in the speed range 724–1656 min1 are plotted in Fig. 5.77a.
For the HP pump belonging to the same RO feed pump installation, the
characteristics of delivery head, power demand, efficiency, and NPSHr as functions
of pump delivery are shown in Fig. 5.77b.
Infinitely variable speed control of a pump is possible by means of mechanical
variable speed gearing as well as by varying the speed of its drive motor. Of these
two alternatives, stepless control of motor speed by frequency conversion is the most
commonly applied method in process technology today.
The synchronous speed nS of a drive motor depends on the frequency fC of the
input current and the number of poles Np of the motor (Eq. 5.225). For a synchronous
638 5 Reverse Osmosis Membrane System: Core Process of SWRO

motor, the synchronous speed nS also corresponds to the speed nM of the rotor,
i.e. the motor. To determine the speed of an asynchronous/induction motor at full
load, its slip rate must be taken into account (Eqs. 5.225a and 5.225b). The slip rate
depends on the rated power of the motor and decreases as the motor’s power output
increases. The slip rate can vary between 0.5% and 5%.

Fig. 5.77 (a) RO supply/booster variable speed pump with operating characteristics plotted
against speed (Courtesy: #KSB SE & Co. KGaA). (b) RO system high-pressure pump
characteristics graph (Courtesy: #KSB SE & Co. KGaA)
5.5 RO Process Units, Components and Overall System Configuration 639

Fig. 5.77 (continued)

fC
nS ¼ 120  : ð5:225Þ
Np
 
s
nM ¼ nS  1  : ð5:225aÞ
100
f  
s
nM ¼ 120  C  1  : ð5:225bÞ
Np 100

nS ¼ synchronous speed of motor


fC ¼ frequency of input current [Hz]
Np ¼ number of poles of motor
640 5 Reverse Osmosis Membrane System: Core Process of SWRO

nM ¼ speed of motor [min1]


s ¼ slip rate of asynchronous motor [%]
For a 4-pole motor, the synchronous speed is 1500 min1 at a mains frequency of
50 Hz. If a motor is controlled by a frequency converter, its speed is thus reduced in
proportion to the reduction in frequency.
The cost of the electrical and electronic devices that make up a frequency
converter depends on the rated power of the drive motor and increases steeply as
the rated power of the motor increases. Hence, it is not only process and control
engineering considerations but also cost aspects that explain why, in an RO feed
pump installation, power and pressure are varied by speed control mainly of the
supply pumps with their lower motor power rating and less at the HP pumps.

5.5.2.2.3 Energy Recovery Systems


Energy recovery systems exploit the residual pressure in the RO concentrate after it
has passed through the membrane system by transferring it to the feed side to reduce
the energy required to raise the feed pressure to that required for passage of the
seawater through the RO system.

Concentrate Pressure Available for Energy Recovery


The concentrate pressure pC(d ) is the result of the delivery pressure of the HP pump
pP, T(d ), which is the feed pressure pF,T,RO,ðdÞ,ΔτD to the RO unit minus the pressure
loss in the flow channels of the membrane elements ΔpF,CðdÞ,ΔτD as shown in
Eq. (5.215j), the pressure loss in the pressure vessels ΔpPV(d ), and the pressure
losses both in the piping to the membrane system and within this Δppiping(d ).
Depending on the type of pressure vessel selected and its configuration (see Sect.
5.5.2.1), its pressure loss is obtained by consulting the suppliers’ documentation.
When calculating the concentrate pressure from the pressure head of the HP pump,
the height difference within the feed pump installation up to the membranes must
also be taken into account as a pressure difference Δpgeo (Eq. 5.226).
 
pC ð d Þ ¼ pP,T ðdÞ  ΔpF,CðdÞ,ΔτD þ ΔpPVðdÞ þ ΔppipingðdÞ þ Δpgeo
 
¼ pF,T,ROðdÞ,ΔτD  ΔpF,CðdÞ,ΔτD þ ΔpPVðdÞ þ ΔppipingðdÞ ð5:226Þ

pF,T,ROðdÞ,ΔτD þ ΔpM,C,RO:

pC(d ) ¼ concentrate pressure at design conditions [bar]


pP, T(d ) ¼ HP pump pressure at design conditions [bar]
ΔpF,CðdÞ,ΔτD ¼ pressure loss of feed to concentrate of flow channels of fouled
membranes at design conditions (d ) and after deposition time ΔτD [bar]
ΔpPV(d ) ¼ pressure loss within the pressure vessels [bar]
Δppiping(d ) ¼ pressure loss within piping and valves at design conditions [bar]
Δpgeo ¼ pressure difference from geodetic difference from the hydraulic profile
[bar]
5.5 RO Process Units, Components and Overall System Configuration 641

pF,T,ROðdÞ,ΔτD ¼ feed pressure needed for fouled RO unit at design conditions [bar]
ΔpM, C, RO ¼ summation of pressure losses of RO unit [bar]
In a reverse osmosis system which, as is usually the case in seawater desalination,
is operated at a constant product flow rate FPr, the pressure losses in the membrane
system are compensated by a corresponding proportional increase in the RO feed
pressure. In this case, Eqs. (5.226) and (5.215i) provide the relationship according to
Eq. (5.227), namely that the concentrate pressure pC(d ) corresponds to the feed
pressure pF, M(d ), clean needed for the “clean” membrane plus the pressure increase
to compensate for the CEOP effect Δπ CEOP,ΔτD and, if an isobaric energy recovery
system is used, the increase in the salinity ΔpF, RO(d ), mix of the feed to the
membranes.

pC ð d Þ ¼ pF,T,ROðdÞ,clean  ΔpFCðdÞclean þ ΔpF,ROðdÞ,mix þ Δπ CEOP,ΔτD


¼ pF,ROðdÞ,clean þ ΔpF,ROðdÞ,mix þ Δπ CEOP,ΔτD pF,ROðdÞ,clean þ ΔpFROðdÞ,mix :
ð5:227Þ

pF, T, RO(d ), clean ¼ feed pressure total needed for the clean membrane system at
design conditions ¼ membrane feed pressure output by membrane manufacturers’
design software [bar]
ΔpFCðdÞclean ¼ pressure loss of clean flow channels ¼ output of membrane
manufacturers’ design software [bar]
pF, RO(d ), clean ¼ feed pressure needed for clean membrane system without flow
channel loss ¼ output of membrane manufacturers’ design software [bar]
ΔpF, RO(d ), mix ¼ pressure increase at design conditions due to rise in salinity in
the RO feed if isobaric ERD systems are used [bar]
Δπ CEOP,ΔτD ¼ CEOP induced increase of osmotic pressure [bar]

Commercialized Energy Recovery Systems


For recovering energy from the concentrate of RO seawater desalination processes, a
number of commercial energy recovery device (ERD) systems are available, which
differ in their technical design and their efficiency. Their mode of operation can be
divided into two categories on the basis of the energy recovery principle, namely
whether the pressure or the energy of the concentrate is transmitted:

• to centrifugal systems driven by the concentrate and directly coupled to the HP


pump side, where they reduce the pressure to be supplied by the HP pump or the
energy required to supply it. These ERD systems include the Pelton turbine from
Calder-Flowserve and the turbocharger systems offered by Energy Recovery, Inc.
(ERI) and FEDCO.
• to systems in which the transferred energy is used to generate pressure in a
separate fraction of the RO feed stream and this part stream is combined with
the fraction from the RO HP pump installation in the feed line to its membrane
system to make up the entire RO feed stream, termed isobaric systems. Such
642 5 Reverse Osmosis Membrane System: Core Process of SWRO

ERD systems are referred to as work exchangers, e.g. the DWEER system from
Calder-Flowserve and the Saltec system from KSB, as well as the pressure
exchanger (PX) from ERI.

Pelton Turbine
The Pelton turbine (PT) ERD system consists of a rotor with spoon-shaped buckets
(Fig. 5.78). The concentrate from the RO unit is directed as a high-pressure jet to the
buckets of the turbine rotor through injector nozzles, causing it to rotate. This
rotational energy is transferred via the Pelton turbine shaft to the drive motor of
the HP pump unit, so reducing the energy from the motor that is required to drive it.
Other methods of energy transmission from the Pelton turbine to the HP pump
unit are also possible. However, mechanical coupling of the Pelton turbine directly
to the drive motor of the HP pump, as shown in Fig. 5.79, is the most common
configuration of this ERD system that is installed in industrial-scale RO plants.
Depending on its rating, the energy recovery turbine (ERT) is equipped with one
or two runners with their associated injector nozzles of which there could be from
one to four, likewise depending on the rating of the turbine. The concentrate is
completely depressurized to ambient pressure in the ERT and must also be
discharged from the turbine casing without pressure.

Fig. 5.78 Energy recovery Pelton turbine (Image provided courtesy of Flowserve US, Inc., all
rights reserved)
5.5 RO Process Units, Components and Overall System Configuration 643

Pelton turbine High pressure pump

Motor

Fig. 5.79 Pelton turbine—motor—HP pump configuration (Image provided courtesy of


Flowserve US, Inc., all rights reserved)

High pressure
pump HP ∆pM,C,RO

pF,HP Pp,T ∆pF,RO pF,T,RO FPr


RO
FF,HP
∆pF,SP
M
pp,SP ~ FC pC
Pelton
turbine PT ∆pC
VSD pC,PT
SW supply
pump SP

FSW,in pC,out
pSW,in
FC,out

Fig. 5.80 Pelton turbine ERD system—flow schematic

How a Pelton turbine is integrated into an RO plant for energy recovery is shown
in the flow schematic of Fig. 5.80.
For the design parameters shown in Fig. 5.80, the power requirement PD, PT of the
core RO system, comprising an RO array with HP pump and coupled ERT, is
calculated according to Eq. (5.228).
644 5 Reverse Osmosis Membrane System: Core Process of SWRO

F SW,in ðppT pF,HP Þ


F C ðpC,PT pC,out ÞηPt

36ηHP 36
PD,PT =
ηM,HP
   
F SW,in  ppT  pF,HP  F C  pC,PT  pC,out  ηPt  ηHP
¼ : ð5:228Þ
36  ηHP  ηM,HP

PD, PT ¼ power demand with Pelton turbine as ERT [kW]


FSW, in ¼ seawater feed flow to RO [m3/h]
ppT ¼ HP pump delivery pressure to RO feed piping system [bar]
pF, HP ¼ pressure of supply to HP pump [bar]
ηPt ¼ efficiency factor of Pelton turbine []
ηHP ¼ efficiency factor of HP pump []
ηM, HP ¼ efficiency factor of HP pump motor []
FC ¼ concentrate flow to Pelton turbine [m3/h]
pC, PT ¼ concentrate pressure to Pelton turbine [bar]
pC, out ¼ Pelton turbine concentrate-discharge pressure ¼ 0 for Pelton turbine
[bar]
Referred to the product water flow FPr and the RO recovery rate YRO, the power
demand PD, PTis determined as shown in Eq. (5.228a).

F Pr
PD,PT =
36  Y RO
"   #
ppT  pF,HP  ð1 2 Y RO Þ  pC,PT  pC,out  ηPt  ηHP
 : ð5:228aÞ
ηHP  ηM,HP

FPr ¼ product water flow [m3/h]


If the feed pump installation is additionally equipped with a supply pump and this
is included in the calculation of the RO unit’s power requirement, Eq. (5.228b)
together with the design parameters shown in the flow schematic of Fig. 5.80 apply.

F Pr
PD,PT ¼
36  Y RO
"   #
pp,SP  pSW,in ppT  pF,HP  ð1 2 Y RO Þ  pC,PT  pC,out  ηPt  ηHP
 þ
ηSP  ηM,SP  ηVSD,SP ηHP  ηM,HP
ð5:228bÞ

pp, SP ¼ supply pump pressure [bar]


pSW, in ¼ pressure of seawater feed to supply pump [bar]
ηSP ¼ efficiency factor of supply pump []
ηM, SP ¼ efficiency factor of motor supply pump []
ηVSD, SP ¼ efficiency factor of variable speed drive supply pump []
5.5 RO Process Units, Components and Overall System Configuration 645

The active residual pressure pC, PT in the RO concentrate for energy recovery as
feed pressure to the Pelton turbine—ERT is calculated with Eqs. (5.228c) and
(5.228d) or, if the RO unit is operated at constant product flow, with Eq. (5.228e).
 
pC,PT ¼ pF,T,RO  ΔpM,C,RO þ ΔpC : ð5:228cÞ

ΔpM,C,RO ¼ Δp f ,F,C,M ðdÞ,ΔτD þ ΔpPVðdÞ þ ΔppipingðdÞ : ð5:228dÞ


 
pC,PT ¼ pF,T,M ðdÞ,clean þ Δπ CEOP,ΔτD  ΔpFCðdÞclean þ ΔpC
  ð5:228eÞ
pF,T,M ðdÞ,clean  ΔpFCðdÞclean þ ΔpC :

pF, T, RO ¼ pF,T,ROðdÞ,ΔτD ¼ necessary feed pressure to fouled RO unit at design


conditions [bar]
ΔpM, C, RO ¼ pressure loss in RO [bar]
ΔpC ¼ pressure loss in concentrate piping [bar]
The inlet pressure parameters for the supply pump and the HP pump as well as the
feed pressure to the membrane system are calculated with Eqs. (5.228f), (5.228g),
and (5.228h).

pp,T ¼ pF,T,RO þ ΔpF,RO : ð5:228fÞ

pF,T,RO ¼ pp,T  ΔpF,RO : ð5:228gÞ

pF,HP ¼ pp,SP  ΔpF,SP : ð5:228hÞ

ΔpF, RO ¼ pressure loss in RO feed piping [bar]


ΔpF, SP ¼ pressure loss in supply pump feed piping [bar]
If, after being relieved to atmospheric pressure in the Pelton turbine, the RO
concentrate cannot be discharged under gravity, i.e. if a height difference above the
ERT level has to be overcome to its discharge point, the pumping power required for
this must additionally be taken into account in the energy calculation of the RO unit
according to Eq. (5.228i).

F C  ΔH out  ρC
PD,PT,dis =
3:671  102  ηPdis  ηM,Pdis
F Pr  ð1 2 Y RO Þ  ΔH out  ρC
¼ : ð5:228iÞ
3:671  102  ηPdis  ηM,Pdis  Y RO

PD, PT, dis ¼ power demand for RO concentrate-discharge pumping [kW]


ΔHout ¼ pressure head increase needed for discharge [m]
ρC ¼ density of concentrate [kg/m3]
ηPdis ¼ efficiency factor of discharge pump []
ηM,Pdis ¼ efficiency factor of motor discharge pump []
646 5 Reverse Osmosis Membrane System: Core Process of SWRO

The specific total energy consumption of the RO unit SECPT, T referred to the
product flow FPr is calculated with Eq. (5.228j) as the quotient of the power
requirement of the core system PD, PT, that is the core system with supply pump
PD, PT plus any additional power requirement PD, PT, dis required to transport the
concentrate to the installation’s discharge point, and the product flow FPr.

PD,PT þ PD,PT,dis
SECPT,T = : ð5:228jÞ
F Pr
SECPT, T ¼ total specific energy consumption with Pelton turbine [kWh/m3]
The Pelton turbine can attain an efficiency of close to 89%, i.e. an efficiency
factor ηPt of 0.89, although this applies only for its design point, i.e. its BEP. Should
the concentrate feed rate or its feed pressure to the turbine change, this also
influences its efficiency and, depending on how much these two parameters change,
this can also result in an efficiency reduction. This behaviour is shown in Fig. 5.81 in
a Hill Chart for a Pelton turbine rotating at 3000 min1 and a design pressure head of
680 m as a function of the concentrate flow rate and the pressure head. Such changes
in the concentrate’s pressure head and feed flow occur in the operating behaviour of
an RO unit if its recovery rate YRO is reduced or increased. If YRO is reduced,
concentrate flow increases and the inlet pressure to the ERT decreases; if the

Fig. 5.81 Pelton turbine—hill chart (Image provided courtesy of Flowserve US, Inc., all rights
reserved)
5.5 RO Process Units, Components and Overall System Configuration 647

recovery rate increases, concentrate flow decreases and the inlet pressure increases.
The turbine efficiency changes accordingly.
However, as Eq. (5.228) shows, the efficiency of the complete system consisting
of the HP pump and the Pelton turbine connected to it depends on the efficiency of
both the turbine and the pump.
Pelton turbines for energy recovery are available for a delivery range from 15 to
1200 m3/h of concentrate; this corresponding to production capacities at a recovery
rate YRO of 40% from 240 to 19,200 m3/day. Thus, they can be used for both smaller
desalination plants and large-scale seawater desalination plants.

Turbocharger
A turbocharger (TC) unit is made up of a turbine impeller, onto which the HP
concentrate flow from the RO unit is directed, and a pump impeller, which is rigidly
connected to the turbine impellor via a common shaft. The turbine is set in rotation
by the pressure power of the concentrate and this drives the pump part of the
turbocharger via the shaft. While the concentrate is depressurized due to transfer
of its energy to the turbine rotor, the feed flow into the pump is brought to a
correspondingly higher pressure (Figs. 5.82 and 5.83).
According to Eq. (5.229), the increase in pump pressure or boost pressure ΔpF,
TC that can thus be attained depends on the ratio of the concentrate feed rate FC to the
turbocharger to the rate of feed FSW, in to its pump part, the pressure differential
available in the turbine for pressure relief, and the turbocharger efficiency ηTC.

Fig. 5.82 Cut-away view of turbocharger


648 5 Reverse Osmosis Membrane System: Core Process of SWRO

High pressure Turbocharger


pump HP TC
∆pM,C,RO
∆ pF,TC
pF,HP pp,T pF,T,RO FPr

∆pF,RO RO
∆pF,SP FF,HP pC,out
pp,SP
FC,out
VSD FC
SW supply pC,TC pC
pump SP ∆pC
FSW,in
pSW,in

Fig. 5.83 Turbocharger flow schematic

FC  
ΔpF,TC = ηTC   pC,TC  pC,out : ð5:229Þ
FSW,in

ΔpF, TC ¼ boost pressure of turbocharger [bar]


pC, TC ¼ concentrate pressure to turbocharger [bar]
ηTC ¼ efficiency factor turbocharger []
For energy recovery in a seawater RO system, the turbocharger is usually
employed as a booster pump. For this purpose, it is installed downstream of the
HP pump and takes over part of the pressure increase in the feed line to the
membrane system (see Fig. 5.83). Depending on the boost pressure ΔpF, TC that
can be provided by the turbocharger according to Eq. (5.229), the delivery pressure
for which the HP pump has to be dimensioned is reduced accordingly.
The delivery pressure of the HP pump pp, T that is then required is calculated from
the RO feed pressure pF, T, RO, the pressure loss due to piping and fittings in the HP
feed line to the RO unit ΔpF, RO, and the boost pressure provided by the turbocharger
ΔpF, TC (Eq. 5.229a).

pp,T ¼ pF,T,RO þ ΔpF,RO  ΔpF,TC ð5:229aÞ

pp, T ¼ feed pressure of HP pump [bar]


ΔpF, RO ¼ pressure loss in RO feed line [bar]
With a turbocharger installed in an RO unit, the quotient of the concentrate inflow
rate to the turbine impeller and the inflow rate to the pump section of the turbo-
charger F FSW,in
C
can also be expressed by the RO unit’s recovery rate according to
Eq. (5.229b). Equation (5.229) is then transformed to Eq. (5.229c).

FC
¼ 1  Y RO : ð5:229bÞ
F SW,in
5.5 RO Process Units, Components and Overall System Configuration 649

 
ΔpF,TC ¼ ηTC  ð1  Y RO Þ  pC,TC  pC,out ð5:229cÞ

The inlet pressure of the concentrate to the turbocharger pC, TC is calculated with
Eq. (5.229d) from the RO feed pressure pF, T, RO and the pressure loss in its
membrane system with Eq. (5.228d).
 
pC,TC ¼ pF,T,RO  ΔpM,C,RO þ ΔpC : ð5:229dÞ

According to the manufacturer, the pressure pC, out to which the concentrate
should be relieved should be at least 1 bar. If to then discharge the concentrate can
not be discharged under gravity, with this ERD it is also possible to relieve the
pressure to more than 1 bar so that any height difference or backpressure in the
concentrate-discharge line can be overcome without an additional pumping station.
However, the resulting lower pressure differential available to act on the turbo-
charger impeller results in an increase in the power requirement of the RO feed pump
installation.
The efficiency of the turbocharger ηTC depends on the flow rate of the concentrate
FC admitted to it. Also, the efficiency factor increases with increasing concentrate
flow. Figure 5.84 shows an example of this dependency for the ERD unit of a
turbocharger manufacturer [135].

Turbocharger efficiency
[%]
85

80

75

70

65

60
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800
Concentrate flow [m3/h]

Fig. 5.84 Turbocharger: dependence of efficiency on concentrate feed flow


650 5 Reverse Osmosis Membrane System: Core Process of SWRO

The power demand of the RO core system, that is the HP pump, turbocharger, and
membrane system, is calculated according to Eqs. (5.229e) and 5.229f.
 
F SW,in  pF,T,RO þ ΔpF,RO  ΔpF,TC  pF,HP
PD,TC = : ð5:229eÞ
36  ηHP  ηM,HP
h   i
FC
F SW,in  pF,T,RO þ ΔpF,RO  ηTC  FSW,in  pC,TC  pC,out  pF,HP
PD,TC = :
36  ηHP  ηM,HP
ð5:229fÞ

PD, TC ¼ power demand with turbocharger as ERT [kW]


If the power requirement is referred to the product flow rate FPr and the RO unit’s
recovery rate YRO, it is calculated with Eq. (5.229g).
   
F Pr  pF,T,RO þ ΔpF,RO  ηTC  ð1  Y RO Þ  pC,TC  pC,out  pF,HP
PD,TC = :
36  ηHP  ηM,HP  Y RO
ð5:229gÞ

If the feed pump station also contains a supply pump, the calculation is expanded
as shown by Eq. (5.229h).

F Pr pp,SP  pSW,in
PD,TC ¼ 3
36  Y RO ηSP  ηM,SP  ηVSD,SP
 
pF,T,RO þ ΔpF,RO  ηTC  ð1  Y RO Þ  pC,TC  pC,out  pF,HP
þ : ð5:229hÞ
ηHP  ηM,HP

When referred to the system’s product flow, the specific power requirement of the
RO unit with a turbocharger as ERD is calculated with Eq. (5.229i).

PD,TC
SECTC = : ð5:229iÞ
F Pr
SECTC ¼ specific energy consumption with turbocharger [kWh/m3]
Turbocharger ERDs as used for seawater desalination are available from
manufacturers over a wide delivery range, from less than 100 to over 3000 m3/
h seawater feed capacity. Turbochargers can therefore be used for small and
medium-sized desalination plants as well as for large-scale seawater desalination
plants. At the maximum turbocharger feed rates and for a recovery rate of 40%, a
product flow of up to 30,000 m3/day can be attained.

Isobaric Energy Recovery Systems


In isobaric energy recovery systems, the feed flow to the RO system is divided into
two part-streams. One of these streams is routed via the HP pump, while the other is
5.5 RO Process Units, Components and Overall System Configuration 651

piped to the intake of the energy recovery system. This second part stream exploits
the residual pressure of the RO concentrate by transferring it to the seawater feed of
the membrane system.

Work Exchanger With work exchanger (WX) ERDs, pressure transfer from the
concentrate to the feed part stream takes place in usually two tubular pressure
vessels, or pressure pipes, within which pistons traverse. The operating principle
of the work exchanger is shown in Fig. 5.85.
The two ERD pressure vessels are alternately supplied with seawater and with
concentrate, and the pistons traversing within them serve both to partition the two
solutions with their differing salinities and to transfer pressure. One of the two
pressure vessels, which already contain seawater, is charged with the high-pressure
concentrate and the seawater is brought to the pressure of the concentrate by the
piston movement within it, caused by the supply of the concentrate. At the same
time, the pressure of the concentrate in the second pressure vessel is relieved and it is
displaced by low-pressure seawater admitted to and acting on the other side of the
piston so that it can be discharged from the ERD system. The alternating admission
of pressurized concentrate to the two pressure vessels, the supply of pressurized
seawater to the RO feed line, and the discharge of depressurized concentrate out of
the ERD are controlled by special types of control valves on the seawater and
concentrate part of the work exchanger. Depending on the valve setting during the
different phases of the work exchanger operation, these valves direct concentrate or
seawater into the two pressure vessels of the Work Exchanger.
Due to the pressure losses in the RO membrane system, the concentrate line to the
energy recovery system and within the ERD itself, the seawater pressure at the ERD
high-pressure discharge point is still less than the required RO feed pressure. This
difference in pressure is compensated for by a booster pump installed in the ERD

Seawater Concentrate
LP feed / Control HP feed /
HP RO feed system LP discharge
RO control valves control valve
HP feed
Pressure RO HP
vessels Concentrate feed

Vessel pistons

LP seawater feed HP RO concentrate


Seawater RO LP
HP RO feed LP RO concentrate Concentrate
LP feed
discharge

Fig. 5.85 Work exchanger—operating principle


652 5 Reverse Osmosis Membrane System: Core Process of SWRO

high-pressure discharge line, whose pressure head is calculated depending on the


total of the above pressure losses.
The construction of a DWEER (dual work exchanger energy recovery device)
work exchanger device manufactured by Flowserve—Calder, an isobaric ERD used
in a large number of industrial seawater desalination plants for energy recovery from
RO concentrate, is shown in Fig. 5.86. The two pressure pipes have a standard length
of 7.5 m, excluding the dimensions of the check valves and the LINX control valve,
and an internal diameter of approx. 350 mm. The pistons reciprocate in the tubes at a
maximum frequency of 4–5 cycles per minute. Similar to a piston pump, this
generates a fluctuating discharge flow. When installing and operating this system,
it is therefore important to ensure that the generated pressure pulses will not be
transmitted to the upstream and downstream RO components or, if so, only if they
are mostly damped out.
The pressure pipes, the control valve, and the check valves are made of highly
corrosion-resistant duplex stainless steels, the pressure pipes optionally also of glass
fibre reinforced plastic (GRP) (see Chap. 6 in Volume 2).
The construction of KSB’s work exchanger, the Saltec system, is comparable to
that of Flowserve-Calder’s DWEER system. This ERD, too, consists of two pressure
pipes, although the control valve has rotating control components with the supply of
concentrate and seawater to the work exchanger pipes regulated by their angular
position as they rotate.
These two work exchanger systems can handle concentrate up to a flow rate of
100–350 m3/h for the DWEER system and 250 m3/h for the Saltec system. For a

Fig. 5.86 Work exchanger—DWEER unit (Image provided courtesy of Flowserve US, Inc., all
rights reserved)
5.5 RO Process Units, Components and Overall System Configuration 653

recovery rate, YRO, of 40%, this corresponds to a maximum product flow ranging
from 4000 to 5600 m3/day.
For larger RO systems for which the product flow of the entire plant or its RO
trains is greater than the available capacity of a single work exchanger unit, ERD
modules are assembled in which more than one DWEER unit is connected in
parallel. Such a module with three DWEER units for a product flow of 13,600 m3/
day is shown in Fig. 5.87. In this module, the DWEER units are installed horizon-
tally. Depending on the configuration of the RO unit, a module can also consist of
vertically mounted ERD elements or alternatively DWEER elements can be installed
in an RO rack below the membrane modules. Thus, this type of energy recovery
system can cover a range of seawater desalination capacities from medium-sized up
to large-scale desalination plants with a capacity of several 100,000 m3/day.

Pressure Exchanger With the PX pressure exchanger supplied by Energy Recov-


ery Inc. (ERI), pressure exchange between the high-pressure concentrate and the
low-pressure seawater feed section is performed in a ducted ceramic rotor rotating in
a ceramic sleeve. Two fixed end covers or seal plates, which are provided with
correspondingly profiled openings in the low-pressure and high-pressure sections of
the PX unit, alternately direct concentrate and seawater into the chambers of the rotor
as it rotates (Fig. 5.88). The rotation of the ceramic rotor is generated by the flow of
concentrate and seawater admitted to the PX unit. There is a narrow gap between the

Fig. 5.87 Work exchanger—module with horizontal triple DWEER arrangement (Image provided
courtesy of Flowserve US, Inc., all rights reserved)
654 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.88 PX—cut-away view and core parts (Images provided courtesy Energy Recovery, Inc.
# 2021, All rights reserved)

stationary ceramic sleeve and the moving rotor which is filled with concentrate to
form a hydrodynamic bearing.
The phases of pressure exchange between HP concentrate and the seawater inlet
to the PX-ERD are shown by way of example in Fig. 5.89 within one of the rotor
chambers for one rotor revolution
During the first stage of pressure transfer from the RO concentrate to the feed
seawater, the concentrate that is in the rotor chamber is displaced by the low-pressure
seawater, is depressurized, and flows to the ERD’s concentrate outlet port. Once the
rotor chamber is filled with low-pressure water, in the next stage it is closed and
sealed off at its inlet and outlet ends. Then, from the pressure exchanger’s high-
pressure side, concentrate flows into the chamber where it compresses the
low-pressure feed seawater and forces it to the PX’s high-pressure discharge port.
In the final stage, the chamber is then filled under pressure with concentrate and
closed off and sealed again, and the cycle starts anew with stage 1 and displacement
5.5 RO Process Units, Components and Overall System Configuration 655

Fig. 5.89 PX operating principle (Images provided courtesy Energy Recovery, Inc. # 2021, All
rights reserved)

of the concentrate by the feed seawater. As Fig. 5.89 shows, the seawater and
concentrate are not separated by a physical barrier, as is the case with the work
exchanger system with piston. Thus, in the pressure exchanger system, in the contact
zone between seawater and concentrate, the two media partially mix.
The ceramic rotor rotates at up to 800–1500 min1, depending on the PX type.
Accordingly, at these speeds the noise level of the pressure exchanger system is
relatively high and may be as much as up to above 80 dBA.
The components of the above described pressure exchanger together with the
inlet and outlet ports for the low and high-pressure ends of the PX and the associated
housing end plates are mounted in a GRP pressure vessel. The so-configured PX
elements are available for concentrate flows of 4.5–68 m3/h, corresponding to
product flows of 72–1088 m3/day at a YRO of 40%. The housings of the biggest
PX units are 800–1000 mm long and 300 mm in diameter.
656 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.90 PX module of an RO train (Source: Author)

Like with the work exchanger, this ERD type also caters for greater product flows
from desalination systems or their trains by connecting the PX units in parallel to
make up larger modules. With the pressure exchanger system as described, it is
therefore possible to equip RO seawater desalination plants with energy recovery
systems at the small and very small bottom end of the range right up to large-scale
plants.
A PX module for an RO train handling a product flow of 13,350 m3/h is shown in
Fig. 5.90.
Tie-in of an isobaric energy recovery system into an RO seawater desalination
plant is shown in the flow schematic that is labelled with the flow and pressure
parameters needed for calculation of the system, see Fig. 5.91.

Design Calculation of Isobaric ERDs Calculations of work exchanger and pres-


sure exchanger units are both based on the same configuration that is shown as a flow
schematic in Fig. 5.91.
The flow rates of the two part streams into which the feed flow FSW,in into the RO
unit is split with this type of ERD are:

• for the part stream FF,HP that is supplied via the HP pump, approximately equal to
the product flow FPr of the RO unit
• the part stream FF, ERD that is directed to the ERD to serve for energy recovery.
FF, ERD can be calculated with Eq. (5.230) from the product flow FPr and the RO
product recovery rate YRO1.

1
F F,ERD ¼ F Pr  1 : ð5:230Þ
Y RO

F Pr
F SW,in : ð5:230aÞ
Y RO
5.5 RO Process Units, Components and Overall System Configuration 657

High pressure ∆pM,C,RO


pump HP
Pp,T ∆pF,RO pF,T,RO cF,RO FPr cPr
RO
∆ pHP,BP FF,RO
pF,HP VSD
cSW,in
FF,HP ERD booster
∆pF,HP FC,ERD pC
pump BP
cHP,BP pHP,BP cC,ERD ∆pC
FHP,BP pC,ERD
pF,HP,BP
∆pHP,ERD
∆ pF,HP,BP
Isobaric ∆pC,out pC,out
∆pF,ERD pF,ERD
ERD cC,out
cSW,in FF,ERD FC,out
pHPSP pERDSP ∆pLP ,ERD
VSD

HP supply ERD supply


pump HPSP pump ERDSP

cSW,in pSW,in
FSW,in

Fig. 5.91 Isobaric ERD—flow schematic

FF, ERD ¼ seawater feed flow to ERD [m3/h]


FPr ¼ RO product flow [m3/h]
YRO ¼ RO product recovery rate []
For detailed calculation of isobaric ERD units, however, the flows of the two feed
part streams must be modified by applying design parameters that are specific to the
type of ERD system. These parameters are:

• Leakage (also referred to as lubrication flow): This is a reduction of the concen-


trate flow FC, ERD that is available in the ERD unit for pressure compensation by
the part stream FHP, BP that is recirculated to the RO feed line. It arises in a work
exchanger ERD due to leaks in its control valve or in the pressure exchanger at the
seals between the high and low-pressure sides of its rotor. This flow loss must be
compensated by a make-up flow from the HP pump. The leakage or lubrication
flow FL is calculated with Eq. (5.231) and the fL or the leakage or lubrication rate
with Eqs. (5.231a) and (5.231b).

F L = F C,ERD  F HP,BP : ð5:231Þ


F C,ERD  F HP,BP F FL
fL ¼ ¼ 1  HP,BP ¼ : ð5:231aÞ
F C,ERD F C,ERD F C,ERD
658 5 Reverse Osmosis Membrane System: Core Process of SWRO

f Lð%Þ ¼ f L  100: ð5:231bÞ

FL ¼ leakage (lubrication) flow (m3/h)


fL ¼ leakage (lubrication) factor ()
fL(%) ¼ leakage (lubrication) rate (%)
FC, ERD ¼ high-pressure concentrate flow to ERD [m3/h]
FHP, BP ¼ high-pressure feed flow from ERD to ERD booster pump [m3/h]

• Mixing and overflush: Within the energy recovery device, concentrate partly
mixes with the fed-in seawater, so the salinity of the ERD’s seawater-side
discharge FHP, BP is higher than that of the seawater itself. After the part stream
from the ERD FHP, BP merges with that of the HP pump FF,HP, this results in an
increase of the salinity of the RO feed FF,RO. The extent of mixing is influenced
first by how effectively the concentrate is separated from the seawater in the ERD,
but also by how much the concentrate is displaced from its chambers or channels
before they are refilled with seawater. The degree of mixing can be reduced when
the concentrate is displaced by seawater by injecting more of the latter into the
ERD at a higher flow rate. This is termed overflush and it is done by raising the
delivery of the respective supply pump. This additional overflush flow Fof to the
ERD’s inlet flow is calculated with Eq. (5.231c) and the overflush factor fof or the
overflush rate, respectively, with Eqs. (5.231d) and (5.231e).

F of ¼ F F,ERD  F HP,BP : ð5:231cÞ


F F,ERD  F HP,BP F F,ERD F of
f of ¼ ¼ 1¼ : ð5:231dÞ
F HP,BP F HP,BP F HP,BP

f of ð%Þ ¼ f of  100: ð5:231eÞ

Fof ¼ overflush flow [m3/h]


fof ¼ overflush factor []
fof(%) ¼ overflush rate [%]
The degree of mixing in the energy recovery device is characterized by the
mixing factor fmix, ERD or the mixing rate fmix, ERD(%). It is calculated with
Eq. (5.231f) from the salinity cHP, BP of the ERD’s high-pressure discharge, that of
the concentrate feed to the ERD cC, ERD, and that of the seawater feed cSW, in.

cHP,BP  cSW,in
f mix,ERD ¼ : ð5:231fÞ
cC,ERD  cSW,in

f mix,ERDð%Þ ¼ f mix,ERD  100:

fmix, ERD ¼ ERD mixing factor []


fmix, ERD(%) ¼ ERD mixing rate [%]
5.5 RO Process Units, Components and Overall System Configuration 659

cHP, BP ¼ salinity of HP feed flow from ERD [mg/l]


cC, ERD ¼ salinity of concentrate HP flow to ERD [mg/l]
cSW, in ¼ salinity of seawater feed [mg/l]
The graph of Fig. 5.92 shows how varying the overflush flow rate fof(%) influences
the mixing rate fmix, ERD(%) for isobaric ERD units, taking as examples the DWEER
work exchanger and the PX pressure exchanger.
The rates of leakage, mixing, and overflush are specific for the various types of
isobaric ERDs and also for the different types within the work exchanger and
pressure exchanger categories.
In step with the rise in leakage and overflow rates that must be additionally
compensated by the feed pump installation, the delivery flows of the supply pumps
and the HP pump increase. If an isobaric system is selected for energy recovery, the
slightly higher values for the treatment capacity of the safety filters must therefore
additionally be taken into account in the design. If the ERD system is operated with
overflush, this additional feed flow must also be taken into account when designing
the pretreatment plant.
Other specific parameters defined by the ERD manufacturers for their units are
the pressure drop in the high-pressure ΔpHP, ERD and low-pressure ΔpLP, ERD
sections of the respective ERD components.

ERD mixing rate [%]


6

4
Pressure exchanger
PX

2
Work exchanger
DWEER
1

0
0 1 2 3 4 5 6
Lead (overflush) flow rate [%]

Fig. 5.92 Work and pressure exchangers—dependence of mixing rate on overflush flow rate (Data
source: Flowserve—Calder & Energy Recovery Inc.)
660 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.26 Specific design parameters of isobaric ERD systems


Isobaric ERD system
Parameter Symbol Unit DWEER PX
Leakage (lubrication) rate fL(%) % <0.2 <0.8–1.2a
Mixing rate of ERD at 0% overflush fmix, ERD(%) <2.5 <5–6a
ERD HP section pressure loss at max. ΔpHP, ERD bar <1.4 <0.8–1.1a
concentrate flow
ERD LP section pressure loss at max. ΔpLP, ERD <2.1 <0.7–1.0a
concentrate flow
a
For large capacity units PX 180–PX Q300

The values of these particular parameters for calculating an energy recovery


system are specified by the ERD manufacturers to match the design conditions
and they provide design tools for this calculation. Their values with their dependence
on flow rate, pressure conditions, and temperature range can be found there and also
in the ERD technical data sheets.
For the two ERD systems, DWEER and PX, Table 5.26 compares, as an example,
the values for the above design parameters as published by the manufacturers of
these components and identified in their calculation tools.
If the values for product flow FPr and recovery rate YRO as well as the leakage rate
fL(%) as shown in Table 5.26 are known, the flow rates in the part streams of the
isobaric ERD system are calculated using Eqs. (5.232)–(5.232g) [136] (see
Fig. 5.91). When operating the system with overflush, the mixing rate fL(%) results
from the selected overflush rate fof(%) using the values plotted in Fig. 5.92 or the data
of the respective ERD manufacturer.

1
F F,ERD ¼ F Pr   1  ð f of þ 1Þ  ð1  f L Þ: ð5:232Þ
Y RO

1
F F,HP ¼ F Pr  1 þ  1  fL ð5:232aÞ
Y RO

F SW,in ¼ F F,ERD þ F F,HP ¼ F F,RO : ð5:232bÞ


 
1
F SW,in ¼ F Pr   1  ½ð f of þ 1Þ  ð1  f L Þ þ f L  þ 1 : ð5:232cÞ
Y RO

1
F L = F Pr   1  f L: ð5:232dÞ
Y RO

1
F HP,BP ¼ F Pr   1  ð1  f L Þ: ð5:232eÞ
Y RO
5.5 RO Process Units, Components and Overall System Configuration 661

F Pr
F F,RO ¼ F F,HP þ F HP,BP ¼ : ð5:232fÞ
Y RO

1
F C,out ¼ F Pr   1  ð1 þ f of  f of  f L Þ: ð5:232gÞ
Y RO

Due to the mixing in the isobaric ERD unit, the salinity in the feed to the RO
membrane systems cSW, in is greater compared to that of the seawater intake cF, RO. It
is calculated from the seawater salinity cSW, in, the salinity of the concentrate feed to
the ERD cC, ERD, the product salinity cPr, and the RO recovery rate YRO with
Eq. (5.233).

1
cF,RO = Y RO  cPr þ cC,ERD 3 1 : ð5:233Þ
Y RO

cF, RO ¼ salinity of RO feed after admixture of ERD outflow [mg/l]


cPr ¼ salinity in product [mg/l] [g/m3]
The relevant salinity of the product cpr under design conditions can be determined
using the membrane calculation routine of the membrane manufacturer’s software.
The salinities cC, ERD in the ERD’s HP feed FC, ERD and cHP, BP in its HP discharge
FHP, BP are calculated according to Eqs. (5.233a)–(5.233c together with the above
equations for the flow calculation of the ERD part streams, Eqs. (5.232)–(5.232g.
 
F C,out  cC,out 2 F F,ERD  cSW,in þ F HP,BP  cSW,in 3 1  f mix,ERD
cC,ERD = :
F C,ERD  F HP,BP  f mix,ERD
ð5:233aÞ

cSW,in 3 ðF SW,in þ F F,ERD  F HP,BP Þ 2 F Pr  cPr


cC,out = : ð5:233bÞ
F C,out
 
cHP,BP ¼ cSW,in 3 1  f mix,ERD þ cC,ERD 3 f mix,ERD : ð5:233cÞ

cC, out ¼ salinity of LP concentrate discharge from the ERD [mg/l] [g/m3]
This method of calculating the changed feed concentration cF, RO to the RO
process must be done iteratively, since the salinity of the product cPr is influenced by
the change in cF,RO. This also applies to the salinity cC,ERD in the concentrate feed
FC, ERD to the ERD, which in turn determines cHP, BP and then cF, RO with a loop
function.
cF,RO can also be calculated to a good approximation directly from the seawater
salinity cSW, in, the mixing factor fmix, ERD, the leakage (lubrication) factor f L , the
salinity of the product cPr, and the recovery rate YRO of the RO plant with Eq. (5.233d
[137]. However, here the leakage factor f L is not referred to the concentrate flow FC,
ERD, but to the feed rate FF, RO to the RO process.
662 5 Reverse Osmosis Membrane System: Core Process of SWRO

   Y RO  f mix,ERD ð1Y RO  f L Þ
cSW,in 3 1 2 f mix,ERD  1  Y RO  f L 2 cPr 
cF,RO ffi  1Y RO
:
f mix,ERD ð1Y RO  f L Þ
12 1Y RO

ð5:233dÞ
FL F 3 Y RO
fL = = L :
F F,RO F Pr

f L ¼ leakage (lubrication) factor based on feed flow to RO []


The increase in the salinity ΔcF, RO, mix in the feed to the RO process results from
the difference between the calculated feed salinity cF, RO and the seawater salinity in
the feed to the plant cSW, in (Eq. 5.233e). The percentage increase of the salinity in
the feed to the RO process due to mixing ΔcF, RO, mix(%) is calculated using
Eq. (5.233f).

ΔcF,RO,mix = cF,RO 2 cSW,in : ð5:233eÞ


cF,RO 2 cSW,in
ΔcF,RO,mixð%Þ = 3 100: ð5:233fÞ
cSW,in

ΔcF, RO, mix ¼ salinity increase in RO feed due to mixing [mg/l]


ΔcF, RO, mix(%) ¼ salinity increase rate due to mixing [%]
An approximate value for ΔcF, RO, mix(%) can also be determined with
Eq. (5.233g).

ΔcF,RO,mixð%Þ ffi 1:04  Y RO  f mix,ERDð%Þ : ð5:233gÞ

The necessary increase of the feed pressure ΔpF, RO, mix to the RO process,
equivalent to the increase of salinity in the RO feed by mixing ΔcF, RO, mix, is
calculated as the difference between the osmotic pressure of the RO feed π F, RO and
the osmotic pressure of the seawater π SW, in (Eq. 5.234).

ΔpF,RO,mix = π F,RO 2 π SW,in : ð5:234Þ

ΔpF, RO, mix ¼ necessary RO feed pressure increase due to mixing [bar]
π F, RO ¼ osmotic pressure of RO feed [bar]
π SW, in ¼ osmotic pressure of seawater feed [bar]
These osmotic pressures can be calculated as described under Sect. 3.2.3.3, but
can also be obtained using the membrane design software of the membrane manu-
facturer. A good approximation for ΔpF, RO, mix is obtained with Eq. (5.234a) and,
using Eq. (5.234b), also an estimate can be made for the magnitude of the pressure
increase [138].
5.5 RO Process Units, Components and Overall System Configuration 663

ΔpF,RO,mix ffi 1:43  π SW,in  f mix,ERD  Y RO


2  Y RO
  : ð5:234aÞ
2  ð1  Y RO Þ  1  f mix,ERD

ΔpF,RO,mix ffi 1:04  pF,ROðdÞ,clean  Y RO  f mix,ERD : ð5:234bÞ

pF, RO(d ), clean ¼ necessary feed pressure to the clean membrane system without
feed to concentrate pressure loss of clean membrane at design conditions and
without mixing [bar]
With the maximum values for the percentage mixing rate fmix, ERD(%) of 2.5% to
6% as listed in Table 5.26, for isobaric ERDs there results an increase of the salinity
in the RO feed by mixing ΔcF, RO, mix(%) of 0.8% to 3% of the seawater feed salinity
with recovery rates YRO of 30% to 50%. With a seawater salinity π SW, in of 35 g/l and
a recovery rate YRO of 45%, the pressure increase ΔpF, RO, mix equivalent to the
increase in salinity of ΔcF, RO, mix by mixing, is 0.6–1.5 bar. With a higher salinity,
the amount of the increase in salinity and thus the increase in pressure also increases.
With a seawater feed salinityπ SW, in of 45 g/l and a 45% recovery rate, this is
0.8–1.9 bar.
The two part streams FF,HP and FF,ERD entering the HP pump and the RO plant’s
isobaric energy recovery system can be transported via a supply pump PSP and then
split between the two components. This configuration is possible if the supply
pressure to the two part streams is the same (pHPSP ¼ pERDSP ¼ pSP). If though
part of the feed pressure to the RO process is to be generated by the HP supply pump
(pHPSP > pERDSP), two supply pumps would be required.
Depending on the ERD’s configuration, the power requirement of the RO process
PD, IS results from the sum of the power take-ups of the pumps in the isobaric ERD
system according to Eq. (5.235) if just one supply pump is installed, and according to
Eq. (5.235a if the HP pump and the ERD are fed with separate supply pumps.

PD,IS = PD,PSP þ PD,HP þ PD,BP : ð5:235Þ

PD,IS = PD,HPSP þ PD,ERDSP þ PD,HP þ PD,BP : ð5:235aÞ

PD, IS ¼ power demand of isobaric system [kW]


PD, HPSP ¼ power demand of HP supply pump [kW]
PD, ERDSP ¼ power demand of ERD supply pump [kW]
PD, HP ¼ power demand of HP pump [kW]
PD, BP ¼ power demand of ERD booster pump [kW]
PD, PSP ¼ power demand of common supply pump [kW]
Detailed calculation of the power requirement PD, IS on the basis of the delivery of
each pump, their intake and delivery pressures as well as the respective efficiencies
of the pump, drive motor and, if installed, the efficiency of the speed control gear is
then done using Eqs. (5.236) or (5.236a, depending on the configuration of the
supply pumps.
664 5 Reverse Osmosis Membrane System: Core Process of SWRO

0    
1 @ F SW,in  pSP  pSW,in F F,HP  pp,T  pF,HP
PD,IS ¼  þ ð5:236Þ
36 ηPSP  ηM,PSP ηP,HP  ηM,HP  ηVSD,HP

 1
F HP,BP  pp,T  pFHP,BP
þ A:
ηP,BP  ηM,BP  ηVSD,BP

0    
1 @ F F,HP  pHPSP  pSW,in F F,ERD  pERDSP  pSW,in
PD,IS ¼  þ
36 ηHPSP  ηM,HPSP  ηVSD,HPSP ηERDSP  ηM,ERDSP

   1
F F,HP  pp,T  pF,HP F HP,BP  pp,T  pFHP,BP
þ þ A:
ηP,HP  ηM,HP ηP,BP  ηM,BP  ηVSD,BP

ð5:236aÞ

pHPSP ¼ feed pressure of HP supply pump [bar]


pSW, in ¼ pressure of seawater feed to supply pump [bar]
pERDSP ¼ feed pressure of ERD supply pump [bar]
pp, T ¼ feed pressure of HP/ERD booster pump [bar]
pF, HP ¼ supply pressure to HP pump [bar]
pF, HP, BP ¼ supply pressure to ERD booster pump [bar]
pSP ¼ feed pressure of common supply pump [bar]
ηHPSP ¼ pump efficiency factor of HP supply pump []
ηM, HPSP ¼ motor efficiency factor of HP supply pump []
ηVSD, HPSP ¼ variable speed drive efficiency factor of HP supply pump []
ηERDSP ¼ pump efficiency factor of ERD supply pump []
ηM, ERDSP ¼ motor efficiency factor of ERD supply pump []
ηP, HP ¼ pump efficiency factor of HP pump []
ηM, HP ¼ motor efficiency factor of HP pump []
ηVSD, HP ¼ variable speed drive efficiency factor of HP pump []
ηP, BP ¼ pump efficiency factor of ERD booster pump []
ηM, BP ¼ motor efficiency factor of ERD booster pump []
ηVSD, BP ¼ variable speed drive efficiency factor of ERD booster pump []
ηPSP ¼ pump efficiency factor of common supply pump [bar]
ηM, PSP ¼ motor efficiency factor of common supply pump [bar]
The specific energy consumption SECIC referred to the product flow FPr of the
RO process is calculated using Eq. (5.236b).

PD,IS
SECIC = : ð5:236bÞ
F Pr
SECIC¼ specific energy consumption with isobaric ERS [kWh/m3]
The pressure conditions in the RO system and its associated ERD as needed for
calculating the power demand are determined using the system of equations
5.5 RO Process Units, Components and Overall System Configuration 665

Eqs. (5.237a)–(5.237m). The assignment of the pressure parameters listed below to


the part streams and pumps of the reverse osmosis-ERD configuration are shown in
Fig. 5.91.

pP,T ¼ pF,T,RO þ ΔpF,RO : ð5:237aÞ

pF,T,RO ¼ pF,T,RO,SW þ ΔpF,RO,mix : ð5:237bÞ

ΔpM,C,RO ¼ ΔpF,C,ΔτD þ ΔpPV þ Δppiping : ð5:237cÞ

pF,HP ¼ pHPSP  ΔpF,HP : ð5:237dÞ

pHPSP ¼ PF,HP þ ΔpF,HP : ð5:237eÞ

pERDSP ¼ ΔpF,ERD þ ΔpLP,ERD þ pC,out þ ΔpC,out : ð5:237fÞ

pF,HP,BP ¼ pC,ERD 2 ΔpHP,ERD  ΔpF,HP,BP : ð5:237gÞ

pHP,BP ¼ pP,T þ ΔpHP,BP  pF,HP,BP : ð5:237hÞ

pC,ERD = pC  ΔpC : ð5:237iÞ


 
pC,ERD = pP,T  ΔpF,RO þ ΔpM,C,RO þ ΔpC ð5:237jÞ

pF,ERD ¼ pERDSP  ΔpF,ERD : ð5:237kÞ

pC,out = pF,ERD 2 ΔpLP,ERD  ΔpC,out : ð5:237lÞ

pC,out ¼ pERDSP  ΔpF,ERD 2 ΔpLP,ERD  ΔpC,out : ð5:237mÞ

ΔpHP, ERD ¼ pressure loss on HP side [bar]


ΔpLP, ERD ¼ pressure loss on the ERD LP side [bar]
pF, T, RO, SW ¼ RO membrane feed pressure without admixture from ERD [bar]
With the DWEER work exchanger system, the pressure of the RO concentrate
can be relieved to atmospheric by the ERD. With the PX pressure exchanger, a
pressure of at least pC,out of 0.8–1.0 bar must be maintained in the concentrate
discharged from the ERD. However, both systems can also be operated with
backpressure of the ERD concentrate. If concentrate discharge under gravity is not
possible, this ERD, like the turbocharger, can overcome a difference in head or a
backpressure without the need for an additional pumping station. The required
concentrate pressure pC,out is generated by the ERD supply pump ERDSP or the
common supply pump PSP. The then necessary increased delivery pressure pERDSP
for the ERD feed part stream is calculated as shown by Eq. (5.237n).

pERDSP ¼ pC,out þ ΔpC,out þ ΔpF,ERD þ ΔpLP,ERD : ð5:237nÞ

The energy recovery efficiency ηISof an isobaric ERD unit is determined by the
ratio of the total power consumed PISout in the unit to the power supplied to it PISin
666 5 Reverse Osmosis Membrane System: Core Process of SWRO

(Eq. 5.238). The corresponding flow rates of the part streams and their pressures are
calculated using the above equations (Eqs. 5.232–5.232g for the part streams and
Eqs. 5.237a–5.237m for the pressures).
   
PISout F HP,BP  pC,ERD 2 ΔpHP,ERD þ F SW,in  pF,ERD  ΔpLP,ERD
ηIS = = :
PISin F C,ERD  pC,ERD þ F F,ERD  pF,ERD
ð5:238Þ

ηIS ¼ Energy recovery efficiency of isobaric ERD system []


PISout ¼ power expenditure in ERD [kW]
PISin ¼ power input to ERD [kW]
The efficiency of energy recovery in an RO system in which an isobaric ERD is
installed also can be determined from the relationship of the power losses ΔEIS, i, loss
in the system to the power within the RO concentrate according to Eqs. (5.238a)–
5.238f) [138].
P
ΔPIS,i,loss
ηIS = 1 2 =1
PIS,C
P
36 3 ΔPIS,i,loss
      ð5:238aÞ
F Pr Y RO1  1  pF,T,RO  ΔpM,C,RO þ ΔpC
1

F C,ERD  ΔpHP,ERD
ΔPIS,1,loss = : ð5:238bÞ
36  ηP,BP  ηM,BP  ηVSD,BP
 
F F,ERD  ΔpF,ERD þ ΔpLP,ERD þ ΔpC,out þ pC,out
ΔPIS,2,loss = ð5:238cÞ
36  ηERDSP  ηM,ERDSP
 
F HP,BP  ΔpHP,BP þ ΔpF,RO,mix
ΔPIS,3,loss = ð5:238dÞ
36  ηP,BP  ηM,BP  ηVSD,BP
2

F Δp þ ΔpF,RO,mix 1
ΔPIS; 4,loss ¼ Pr 3 4 F,RO þ fL  1 ð5:238eÞ
36 ηP,HP  ηM,HP Y RO
3

pP,T pHP,BP 5:
3 2
ηP,HP  ηM,HP ηP,BP  ηM,BP  ηVSD,BP

F F,HP  ΔpHP
ΔPIS,5,loss = : ð5:238fÞ
36  ηHPSP  ηM,HPSP  ηVSD,HPSP

PIS, C ¼ power available from RO concentrate [kW]


ΔPIS, i, loss ¼ power losses in ERD system [kW]
5.5 RO Process Units, Components and Overall System Configuration 667

The power losses ΔPIS, i, loss result from the pressure losses in the high and
low-pressure sections of the ERD, the pressure increase in the RO feed due to
mixing, the leakage rate of the ERD, the pressure losses of the safety filters
downstream of the supply pumps, and further pressure losses in the pipelines of
the entire system.
Calculation of energy recovery efficienciy rates of reverse osmosis systems for
seawater desalination with isobaric ERD units with these algorithms results in values
of ηIS up to 96%.
Leakage as well as mixing degrade the efficiency of isobaric ERDs. The use of
overflush when operating the ERD can reduce mixing. The extent to which overflush
may be used, however, is established by balancing the increase in the system’s
power requirement due to the pressure increase in the RO supply by mixing against
the increased power requirement of the ERD supply pump for providing the
overflush power [139]. With the work exchanger, the mixing rate can be reduced
primarily by improved sealing of the reciprocating pistons in the unit’s pressure
vessels. In the case of the pressure exchanger, mixing in the contact zone between
concentrate and seawater during the compression phase has to be reduced by
optimizing the flow conditions in the rotor chambers.
The number of ERD units NU,ERD needed to make up an ERD module to attain a
defined performance of an energy recovery system is given by the concentrate flow
FC of the RO unit to be designed, which is equal to FC,ERD or product flow FPr and
recovery rate YRO of the RO system and the design concentrate flow FU, D, C, ERD of
the selected ERD units (see Eq. 5.239).
 
F C,ERD,RO F Pr  Y 1RO  1
N U,ERD ¼ ¼ : ð5:239Þ
F U,D,C,ERD F U,D,C,ERD

NU, ERD ¼ number of ERD units []


FC, ERD, RO ¼ design concentrate flow of RO unit [m3/h]
FU, D, C, , ERD ¼ design concentrate feed flow of ERD unit [m3/h]
Isobaric ERD systems are also used for small- and medium-sized membrane
desalination plants, in which the functions of the ERD and the booster pump,
sometimes even those of the HP pump, are combined in an ERD unit. In some
cases, the positive displacement pump principle is utilized for these functions to
exploit the higher efficiency of this pump type, for example, the iSave system from
Danfoss and the Salino system from KSB.
A configuration comparable to isobaric energy recovery can also be realized with
a turbocharger if it is equipped with an additional drive motor with infinitely variable
speed control to provide additional power to compensate for the respective pressure
losses in the RO unit, like the HP-HEMI system from FEDCO [140].

Dependence of Energy Recovery Efficiency and Specific Energy Consumption


on Product Recovery Rate
In all energy recovery systems described above, the efficiency of energy recovery
and thus also the specific energy consumption of a seawater RO plant are
668 5 Reverse Osmosis Membrane System: Core Process of SWRO

significantly influenced by the overall efficiency of the pumps in the system, i.e. the
pumps themselves, the driving motor, and the speed control unit, as well as the RO
recovery rate.
The graphs in Fig. 5.93 show how the specific energy consumption (SEC) of an
RO plant depends on its recovery rate YRO for the above described energy recovery
systems of Pelton turbine, turbocharger, and isobaric ERD. These values are for a
seawater salinity of 35,000 mg/l at a temperature of 25  C. Other calculation
parameters specific for the system’s energy consumption are stated in the graphic.
For calculating these three systems, a seawater feed pressure of 2 bar to the ERD
was assumed. However, these specific energy consumption figures do not include
energy losses in the RO plant apart from those of the ERD nor the energy
consumptions of the supply pumps. These are figures taken for comparison
purposes, but their values are lower than what would actually be the case for the
design of RO plants in practice. The SEC data shown also only apply if the pumps in
the energy recovery systems are operated at their optimum operating point.
The curves of the diagram of Fig. 5.93 show the differences between the
efficiencies of the various energy recovery systems and the impact of RO recovery
rate on their specific energy consumption. Both the curves of the turbocharger, the
Pelton turbine, and also the isobaric ERD curve have a minimum of the specific
energy consumption SEC at a range of between 38 and 40% of the product recovery
rate YRO1.

Specific energy
consumption SECRO1
[kWh/m3]
2.8

Seawater TDS = 35,000 mg/l


2.7
Temperature = 25°C
Specific membrane flux = 13 l/m2,h
2.6
HP pump efficiency = 87 % Turbocharger
ERD booster pump efficiency = 84%
2.5
Motor efficiency = 96%

2.4

2.3
Pelton turbine
2.2

2.1 Isobaric ERD

2.0

1.9
25 30 35 40 45 50 55
Product recovery YRO1 [%]

Fig. 5.93 Dependence of the SEC of ERD units on RO product recovery rate (data source: Energy
Recovery Inc., Calder—Flowserve)
5.5 RO Process Units, Components and Overall System Configuration 669

Fig. 5.94 Wound filter


cartridge elements (Courtesy:
Rosedale products Inc.)

5.5.2.3 Safety Filters


Safety or guard filters have the task of intercepting any particulate impurities still
present in the RO feed after pretreatment or which have been introduced into the feed
lines as a result of operational faults or during installation work before they can get to
the membranes. If following pretreatment the silt density index SI is less than 3–5,
only a very small amount of particulate matter has to be captured by these filters.
They are therefore normally equipped with non-backwashable cartridge filter
elements that are replaced when a certain differential pressure is attained due to
the build-up of solids (Fig. 5.94).

5.5.2.3.1 Cartridge Filter Elements


According to the specifications of the membrane manufacturers, these safety filter
elements should have an absolute pore size of 5–10 μm. When selecting the
elements, a distinction must be made between nominal and absolute retention
grade. With nominal retention, suspended particulates with a size corresponding to
the nominal filtration degree of the cartridge are retained at a certain percentage, for
instance more than 90–95%, while with absolute retention, 100% retention takes
place at the specified filtration degree.
670 5 Reverse Osmosis Membrane System: Core Process of SWRO

Polypropylene is usually used as the cartridge filter medium material for seawa-
ter. Depending on the filter medium structure, a distinction must be made between:

• wound or spun cartridges: For these, polypropylene fibres are wound around a
supporting core also made of polypropylene with the fibre thickness and type of
winding determining the cartridge’s filtration degree
• melt-blown cartridges: In the melt-blown process, a multi-layer structure with
several microfibre layers of differing porosities is produced from molten plastic
granulate by means of hot-air operated spinnerets with finer and coarser continu-
ous filaments. The filter layer may be deposited onto a support core, but also
cartridges without a support core can be produced.
• pleated cartridges: In these cartridges, the filter medium consists of pleated filter
sheets in which the type of filter medium and its pleating and pleating density
determine the degree of particle retention.

Through appropriate manufacturing processes of these cartridge types, the aim is


not only to intercept particulates on the cartridge surface, but also to achieve
filtration in-depth, so increasing solid absorption capacity and extending service life.
With wound cartridges, filtration takes place from the outside into the cartridge
interior. Depending on the design of the filter medium, with melt-blown and pleated
cartridges, the medium to be filtered can be supplied to the outside or the inside of
the filter.
Depending on the type of filter cartridge, how the medium to be filtered is fed, and
the filtration medium, the cartridges have differing external and internal diameters
and lengths.
If for cartridge filters their primary function is to act as safety filters, more cost-
effective wound cartridges find application as well as melt-blown cartridge filter
elements that are specially designed for this application, with a nominal filtration
degree of 5 μm. If the cartridge filters are also intended to achieve an enhanced
reduction in the silt density index SDI, the filters are fitted with more expensive filter
elements such as pleated cartridges by means of which absolute retention at the
specified pore size is possible.

5.5.2.3.2 Cartridge Filter Housings and Location of Cartridge Filters


A cartridge filtration unit consists of the filter housing and the filter elements, which
are arranged in the housing depending on how the medium is admitted to the
cartridges. The filtration capacity and filtration degree of the cartridge filter are the
factors that determine the type of cartridge assembly, the number of cartridges, and
their length. For smaller capacities, there may be only a few cartridges, which when
they are changed out are individually removed and replaced with new units. For
higher capacities, the cartridges are bundled into packages which when they are
spent are removed from the filter and replaced by new ones. Hoists for filter handling
have to be installed in the filter station for opening the cartridge filters and removing
the cartridge packages.
5.5 RO Process Units, Components and Overall System Configuration 671

Fig. 5.95 Vertical cartridge filter housing

A vertical cartridge filtration housing with the cartridge filter elements arranged in
a removable package and charged with the medium to be filtered from outside to
inside of the cartridges is displayed in Fig. 5.95.
The horizontal design of the filter housing is used mainly for cartridge filters with
a high filtration capacity. As Fig. 5.96 shows, in this case the medium flows through
the cartridges from the inside with the filtrate discharged to the outside.
Safety cartridge filters are normally installed in the feed line of an RO plant. If the
feed pump installation consists of a supply pump and an HP pump, the cartridge filter
is placed between these two pumps. Figure 5.97 shows such a gallery of supply
pumps with their downstream vertical cartridge filters at a large seawater RO plant.
All supply lines of an RO plant must be safeguarded in this way to protect the
membrane system from particulate contamination. This applies likewise to the
chemical membrane cleaning equipment as well as to all chemical dosing stations
672 5 Reverse Osmosis Membrane System: Core Process of SWRO

Fig. 5.96 Horizontal cartridge filter housing

Fig. 5.97 Cartridge filter and supply pump gallery of a large SWRO plant (source: author)

supplying the membranes downstream of the safety cartridge filter station from
which the RO system is fed. These dosing stations must then be equipped with
their own cartridge safety filters.
The cartridge filter housings for seawater filtration may be of rubber-lined steel,
plastic, or stainless steel (see also Chap. 6 in Volume 2).
5.5 RO Process Units, Components and Overall System Configuration 673

5.5.2.3.3 Cartridge Filter Sizing


Cartridge filter elements with the medium flowing from the outside to the inside, as
most commonly used in safety filters, have an outside diameter of 60–64 mm (~2.5
inches), an inside diameter of 2.5–2.8 mm (~1"), and are available in lengths from
254 mm (10") to 1270 mm (50") or 1780 mm (70"), depending on the filter element
manufacturer.
The number of filter elements NCfe required for a defined cartridge filtration
capacity FCf is calculated as the quotient FCf of this filtration capacity and the
filtration capacity FCfe of the specified element (see Eq. 5.240).

F Cf
N Cfe ¼ : ð5:240Þ
F Cfe
NCfe ¼ number of cartridge filter elements []
FCf ¼ feed flow to cartridge filtration installation [m3/h]
FCfe ¼ design feed flow to a single cartridge filter element [m3/h]
FCfe is usually determined from the filtration flow Fs, 10" into a filter element with
a length of 10" and an outside diameter of 2.500 . The usual design values for this
parameter are 0.7–1.2 m3/h. The filtration capacity FCfe of the filter element for its
selected length is then derived from the predetermined filtration capacity Fs, 10" of
the 10" element and the length lCfe of the element using Eq. (5.240a).

lCfe
F Cfe ¼ F s,10}  : ð5:240aÞ
0:254
Fs, 10" ¼ feed flow to selected 10" (254 mm) cartridge filter element [m3/h]
lCfe ¼ length of cartridge filter element [m]
However, FCfe can also be calculated from the filtration velocity vCfe which is
applied to the cartridge filter element and its filtration area SCfe (Eq. 5.240b). If the
filtration flow is from the outside to the inside, the filtration area SCfe of the element is
calculated from its outside diameter dCfe,a and the element length lCfe with
Eq. (5.240c). If filtration is from the inside of the cartridge to the outside, then the
filtration velocity has to be calculated based on the inside diameter dCfe,i of the
cartridge.

F Cfe ¼ vCfe  SCfe : ð5:240bÞ

SCfe ¼ d F,Cfe  π  lCfe : ð5:240cÞ

vCfe ¼ Filtration velocity of cartridge filter element selected [m3/m2, h] [m/h]


dF, Cfe ¼ Diameter of cartridge filter element selected at filtration surface[m]
SCfe ¼ filtration surface area of cartridge filter element [m2]
With the above design values of Fs, 10", the dimensions of the standard 10"
element result in a filtration velocity range of 14–25 m/h.
The criteria for specifying the feed flow Fs, 10" to the 10" element and the filtration
velocity vCfe are:
674 5 Reverse Osmosis Membrane System: Core Process of SWRO

• the pressure loss of the clean cartridge filter element


• whether reserve filters are provided in the cartridge filter station
• the mode of seawater pretreatment.

The pressure loss of an element under the design conditions for current tempera-
ture and feed salinity is calculated from its standard pressure loss ΔpCfe(St) stated for
the assumed filtration flow, taking into account the ratio of dynamic viscosity under
design conditions μ(d ) and that under standard conditions μ(St), as shown in
Eq. (5.240d).

μðd Þ
ΔpCfeðdÞ ¼ ΔpCfeðStÞ  : ð5:240dÞ
μðStÞ

ΔpCfe(d ) ¼ differential pressure of cartridge filter element at design conditions


[bar]
ΔpCfe(St) ¼ differential pressure of cartridge filter element at standard conditions
[bar]
μ(St) ¼ dynamic viscosity at standard conditions (temperature and salinity)
[kg/ms] [cPoise]
μ(d ) ¼ dynamic viscosity at design conditions (temperature and salinity) [kg/ms]
[cPoise]
For calculating the viscosities, see Sects. 3.2.2, 3.2.2.2, Eqs. (3.41), (3.42), and
Fig. 3.16.
For their element types, the cartridge filter manufacturers provide graphs showing
the dependency of the pressure loss ΔpCfe(St) of a 1000 element on its grade of
filtration and feed flow for fresh water at a standard temperature (usually 20  C).
Such a graph is shown in Fig. 5.98 taking as an example a cartridge commonly used
in safety filters.
If the filter station is not equipped with a standby filter for bridging the outage
during a cartridge change at one of the filters, the other operating filters must take
over the additional filtration load of the filter under maintenance. In this case, a
design filtration velocity should be specified more towards the bottom end of the
stated bandwidth of this parameter.
High-flow filter elements allow a significant reduction of the number of elements
for the same filtration flow and filtration degree, as these elements can be operated at
a considerably higher capacity. Such cartridge filter elements are absolute filters, are
equipped with pleated filter media, and the medium flows through them from the
inside to the outside. They have a diameter of 152.4 mm (6") and are available in
lengths from 1016 mm (40") to 1524 mm (60"). They can be supplied for a filtration
velocity of up to 80–100 m/h per element and can be used in both vertical and
horizontal filter housings although horizontal housings are normally preferred, as
with these it is easier to change the filter elements than with vertical housings.
Despite the higher filtration velocity, the pressure loss across these high-flow filter
elements is sometimes the same or even lower than that of conventional cartridges.
5.5 RO Process Units, Components and Overall System Configuration 675

Differential pressure of
cartridge [mbar]
350

Differential pressure of 10 inch (254 mm)


300 cartridge with water at 20°C Filtering grade [micron]
5

250

10
200

150

100

50

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Feed flow [m3/h]

Fig. 5.98 Dependence of pressure loss of 1000 (254 mm) filter cartridge on its feed flow (Data
source: Pall Claris Elements)

High-flow filter elements, though, are considerably more expensive than the
normal wound or melt-blown elements. Thus in order to decide which type of filter
element is to be used for a cartridge filter station, the various cost aspects have to be
compared with regard to differences in dimensions and the number of filter housings
and their cartridge assembly together with filtration degree, pressure loss, and service
life as well as handling of the cartridge packages.
The design of the cartridge filtration station also depends on the extent to which
particulate and colloidal impurities are already removed in the RO pretreatment
stage. If the pretreatment process includes a membrane filtration stage with ultrafil-
tration or microfiltration membranes, the separation efficiency for this type of
impurity would be better there than would be possible with 5–10 μm safety filtration.
Cartridge filtration would then be purely a safeguard without any specific
requirements, such as a reduction in the silt density index. The choice of cartridge
filter elements could accordingly be based on their nominal retention rate and a flow
in the upper range of the filtration velocity may be selected. Often consideration is
given to eliminating safety filtration altogether if the pretreatment stage ahead of the
RO process includes membrane filtration. However, then the RO membranes are no
longer protected in the event of operational malfunctions in the upstream process
stages during which particulate impurities could be generated.
The number of cartridge filter housings NCfh in the safety filter system results
from the number of filter cartridges NCfe needed for the feed flow FCf to be filtered
676 5 Reverse Osmosis Membrane System: Core Process of SWRO

and the number of filter elements NCfe, h with which each filter housing can be
equipped (Eq. 5.240e).

N Cfe
N Cfh ¼ : ð5:240eÞ
N Cfe,h

NCfe ¼ number of cartridge filter elements of filter system []


NCfh ¼ number of cartridge filter housings []
NCfe, h ¼ number of cartridge filter elements per housing []
The filter cartridges must be changed and replaced with new units when the
differential pressure between the inlet and outlet of the cartridge filter reaches a
defined maximum value. Depending on the filtration velocity selected for the
elements and the type and filtration grade of the filter cartridges, the clean cartridges
have values between 0.1 and 0.3 bar. Once the pressure loss of the fouled
cartridges attains 1.2–1.5 bar, they should be replaced.
The total differential pressure ΔpCF(d ) of the cartridge filter results from the
differential pressure ΔpCfe(d ) of the cartridge filter elements themselves plus the
differential pressure ΔpCfh(d ) of the filter housing and associated piping and fittings
for the design or under the prevailing operating conditions (Eq. 5.240f).

ΔpCFðdÞ ¼ ΔpCfeðdÞ þ ΔpCfhðdÞ : ð5:240fÞ

ΔpCF(d ) ¼ differential pressure of cartridge filter at design conditions [bar]


ΔpCfh(d ) ¼ differential pressure of cartridge filter housing and piping at design
conditions [bar]
The maximum pressure loss in the safety filter station of an RO plant will
therefore be about 1.5–2 bar. The time between replacements of the filter cartridges
is determined not only by the build-up of inert solids on the cartridge elements, but
also by the extent of biological growth on the feed side and on the filter itself.
Operating experience with existing plants indicates that, depending on the
pretreatment efficiency and taking into account the design criteria as described
above, the filter cartridges have to be replaced approximately every 3–12 months
of continuous RO operation.
If the filter cartridges have to be changed more frequently than indicated above,
the pretreatment efficiency must be improved with regard to solids separation, but
also the effectiveness of disinfection in the feed line to the filtration equipment and
within the cartridge filter itself must be enhanced.

5.5.2.4 Chemicals Dosing


For conditioning with a combination of acid and antiscalant, these are dosed into the
feed to the main desalination train, whereas for alkaline operation only antiscalant is
dosed at this dosing location. To protect the RO membranes from oxidation, sodium
bisulphite or sodium metabisulphite is added to the RO feed as reducing agent.
If the main RO desalination stage is equipped with cellulose acetate membranes,
due to the higher oxidation resistance of these CA/CTA membranes, chlorine is
5.5 RO Process Units, Components and Overall System Configuration 677

additionally dosed intermittently into their feed line to prevent biological fouling
within the membrane modules. If polyamide brackish water membranes are used in
the second pass, then sulphite must be again dosed into its feed line to remove the
residual chlorine still present in the permeate of the seawater desalination stage.
If polyamide membranes are used in both the main desalination and post-
desalination stages, to raise the pH level caustic soda is added to the second pass
to improve boron retention by the brackish water membranes. In this case, dosing of
antiscalant to the second pass feed is additionally required to prevent scaling in the
post-desalination stage.
Possible locations of the dosing points in the feed lines to the first and second
passes of an RO plant are shown and labelled in Fig. 5.62.

5.5.2.4.1 Acid and Antiscalant Dosing to the RO1 and RO2 Feed: Sizing
and Chemicals Consumption
Acid and Antiscalant Dosing to the RO1 Feed
Dosing of acid and antiscalant in parallel to the feed line of a seawater RO desalina-
tion plant is the most common means of preventing scaling in the concentrate of the
RO1 system. The purpose of acid dosing is to reduce the proportion of bicarbonate
and carbonate present in seawater to such an extent that either the addition of acid
alone prevents calcium carbonate precipitation or, through additional dosing of
antiscalant, there will no longer be any residual tendency to precipitate calcium
carbonate and other scalants (see Sect. 5.3.4.2).
The acid dosing rate RD,Acid,100%,RO1 needed to prevent CaCO3 scaling is calcu-
lated, depending on seawater composition and the RO recovery rate, following the
calculation sequence shown in Table 5.22 in Sect. 5.3.4.2.2. The acid dosing rate
upstream of the RO process depends on the extent to which acid has already been
injected in the pretreatment stage, for example to optimize flocculation, and how this
has changed the composition of the RO feed.
The antiscalant dosing rate RD,antiscale,100%,RO1 for combined acid and antiscalant
dosing is determined by the extent to which the antiscalant must keep both calcium
carbonate and the other seawater scalants in solution.
For the “alkaline” operating principle, there is no acid dosing in the RO feed and
scaling is prevented by antiscalant dosing alone.
For more information on the antiscalant mechanism and calculation of the
antiscalant dosing rate, see Sect. 5.3.4.2.1. A suitable antiscalant for the respective
operating mode and its dosing rate RD,antiscale,100%,RO1 should be identified with the
aid of the software tools provided by the antiscalant manufacturers and in consulta-
tion with them.

Sodium Hydroxide and Antiscalant Dosing to the RO2 Feed


In order to attain the specified boron concentration in the product of a seawater RO
plant for drinking water treatment, the boron rejection rate of the brackish water
membranes of the desalination stage has to be adjusted accordingly. This is done by
raising the pH of the product of the main desalination stage by adding caustic soda to
attain a pH in the range of between 9 and 10.5 in the RO2 feed (see Sect. 5.1.5.2.4).
678 5 Reverse Osmosis Membrane System: Core Process of SWRO

A complex, iterative calculation procedure is required to determine the dosing


rate RD,NaOH,100%,RO2 of caustic soda solution. The simplifying assumptions
described under Sect. 3.2.4.3 for calculating the components of the carbon dioxide
equilibrium and the pH can no longer be applied here. The required pH that has to be
set is outside the range of 6–9 for which the hydroxyl concentration is negligible.
The boron concentration in the product water is also so high in relation to its salinity
that the caustic consumption of the boric acid/borate equilibrium must be taken into
account and the increase in ionic strength due to the addition of caustic soda can no
longer be neglected for exact calculation of the caustic soda dosing rate.
Calculation of RD,NaOH,100%,RO2 is included in the membrane manufacturers’
calculation software. However, a comparison of the values for the dosing rate
calculated with the various software tools shows significant differences, just like
when calculating acid dosing.
The value of RD,NaOH,100%,RO2 can also be calculated with PHREEQC. Due to the
lower salinity of the RO1 product water, for thermodynamic modelling the activity
coefficients of the components involved are no longer calculated with the Pfizer
model, but are determined using the extended Debye-Hückel algorithms such as the
extended Debye-Hückel equation, the Davies equation, or the Truesdell Jones
equation (see also Sect. 3.2.3.2.1). These equation systems find application in the
databases phreeqc.dat, llnl.dat, minteq.dat, minteq.v4.dat, and wateq4f.dat of
PHREEQC, and with one of these PHREEQC software modules, the caustic soda
dosing rate RD,NaOH,100%,RO2 can be calculated.
With increasing pH, the scaling tendency of the scalants still present in the product
water of the main desalination RO1 also increases. These are predominantly alkaline
earth carbonate compounds and the scaling products are dolomite CaMg(CO3)2 and
magnesium carbonate MgCO3 as well as at high pH values magnesium hydroxide
Mg (OH)2 (see Sect. 5.3.4.1.4). For this reason, it is also necessary to dose
antiscalant when injecting caustic soda to increase the pH in the RO2 feed. Here
too, a suitable antiscalant should be selected and its dosing rate RD,antiscale,100%,RO2
calculated using the software tools of the manufacturers of these chemicals and in
consultation with them.

Dosing Systems Sizing and Calculation of Chemicals Consumption


The dosing rate RDC, 100% referred to a 100% concentration of the chemical in
question must be converted to its dosing rate RDC, X% in practice at the concentration
X%del at which it is delivered, as shown in Eq. (5.241).

RDC,100%  100
RDC,X% ¼ : ð5:241Þ
X%del
RDC,100% ¼ dosing rate of chemical at 100% concentration [mg/l ¼ g/m3]
RDC,X% ¼ dosing rate of chemical at delivery concentration X%,del [mg/l ¼ g/m3]
X%del ¼ delivery concentration of chemical [%]
5.5 RO Process Units, Components and Overall System Configuration 679

The amount per hour of the dosed chemical required for a defined flow rate FW at
the delivery concentration X%del is then calculated with Eq. (5.241a) and the annual
consumption of this chemical depending on the RO plant’s operating hours during
this time τo, y is calculated with Eq. (5.241b).

RDC,X%  F W
DC,X% ¼ : ð5:241aÞ
1000
C C,X%,τ0,y ¼ DC,X%  τo,y : ð5:241bÞ

FW ¼ metered water flow [m3/h]


DC,X% ¼ amount of chemical at delivery concentration X%,del [kg/h]
CC,X%,τo,y ¼ consumption of chemical at delivery concentration X%,del during
annual operation time τo, y [kg/year]
τo, y ¼ operation time of RO system per year [h/year]
If the RO plant is operated over 1 year at different dosing rates RDC, X % , i during
certain periods τo, i, the annual consumption of the dosed chemical is calculated as
the sum of the individual consumption values CC,X%,τo,i during these times τo, i
(Eq. 5.241c).

X
i X
i
C C,X%,τ0,y ¼ C C,X%,τ0,i ¼ DC,X%,i  τo,i : ð5:241cÞ
i¼1 i¼1

CC,X%,τo,i ¼ consumption of chemical at delivery concentration of X%,del during


operation time τo, i [kg]
The dosing rate FDC,X% of the chemical dosing equipment is calculated with
Eq. (5.242) on the basis of the flow rate FW of water to be conditioned together with
the dosing rate RDC,X% at the delivery concentration X%del and in case of dosing of
chemical solutions on the basis of their density ρC,X% at the delivery concentration.

RDC,X%  F W RDC,100%100FW
F DC,X% ¼ ¼ : ð5:242Þ
ρC,X%  1000 X%del  ρC,X%  1000

FDC,X% ¼ dosing flow of chemical at delivery concentration of X%,del [l/h]


ρC,X% ¼ density of liquid chemical at delivery concentration of X%,del [kg/l]
So as to be able to dose chemicals at a sufficient rate even under unfavourable RO
plant operating conditions, the calculated flow rate of the dosing devices is
multiplied by a safety factor fsafety. Its value reflects the designer’s assessment of
the need for increased dosing rates for the chemicals in question. The design
injection rate of the dosing device concerned, i.e. the maximum flow FDC, X % ,
max at which the chemical can be dosed, is calculated as shown in Eq. (5.242a).

F DC,X%, max ¼ F DC,X%  f safety : ð5:242aÞ


680 5 Reverse Osmosis Membrane System: Core Process of SWRO

FDC, X % , max ¼ max. dosing flow of chemical at delivery concentration of X%,del


[l/h]
fsafety ¼ safety factor for dosing flow design
Often, though, the chemicals are not dosed at their delivery concentration X%del,
but are first diluted to a dosing concentration of X%dos. Dilution can be carried out
when the dosing solution is made up, but this is often done continuously just before
their injection at the dosing point. For this purpose, the chemical at delivery
concentration is mixed with dilution water, possibly using a static pipeline mixer,
prior to injecting the diluted solution at the dosing point. The dilution water flow rate
required for dilution to the dosing concentration results from the ratio of the delivery
concentration X%del of the chemical solution to its dosing concentration X%dos as well
as the ratio of the density of the solution at delivery concentration ρC, X% to the
density ρDilw of the dilution water (Eq. 5.243).

ρC,X% X%del
F Dilw ¼ F DC,X%   1 : ð5:243Þ
ρDilw X%dos

FDilw ¼ flow of dilution water (l/h, m3/h]


ρC,X% ¼ density of liquid chemical at delivery concentration of X%,del [kg/l]
ρDilw ¼ density of dilution water [kg/l]
X%dos ¼ dosing concentration of chemical [%]
The dosing flow FDC, X % , dos at the dosing point is calculated as the sum of the
dosing flow at delivery concentration FDC,X% and the dilution water flow FDilw
(Eqs. 5.243a–5.243c).

F DC,X%,dos ¼ F DC,X% þ F Dilw : ð5:243aÞ



ρC,X% X%del
F DC,X%,dos ¼ F DC,X%  1 þ  1 : ð5:243bÞ
ρDilw X%dos

RDC,100%100F W ρC,X% X%del
F DC,X%,dos ¼  1þ  1 : ð5:243cÞ
X%del  ρC,X%  1000 ρDilw X%dos

Multiplication by the safety factor fsafety described above fixes the maximum
dosing flow FDC, X % , dos. max that the dosing device in question must provide
according to Eq. (5.243d) and the corresponding flow FDilw, max of dilution water
according to Eq. (5.243e).

ρC,X% X%del
F DC,X%,dos, max ¼ F DC,X%, max  1 þ  1 : ð5:243dÞ
ρDilw X%dos
5.5 RO Process Units, Components and Overall System Configuration 681


ρC,X% X%del
F Dilw, max ¼ f safety  F DC,X%   1 : ð5:243eÞ
ρDilw X%dos

FDC, X % , dos. max ¼ max. dosing flow of chemical at delivery concentration of X%,
del [l/h]
FDilw, max ¼ max. flow of dilution water (l/h, m3/h]
Further details of the calculation of the quantities of chemicals consumed and the
corresponding dosing rates can be found under Sect. 4.2.2.1, and for calculating the
volumes of the storage tanks and dosing vessels under Sect. 4.2.3.2.

5.5.2.4.2 Chlorination and Dechlorination in RO1 and RO2 Feed: Sizing


and Chemicals Consumption
To prevent biological growth in the systems for extracting and pretreating seawater
and also biological fouling of the RO membrane systems, biocide is dosed right into
the desalination plant’s feed line. These biocides could be inorganic and organic
compounds. Chlorine is most frequently used as a disinfectant in RO seawater
desalination in the form of elemental chlorine or chlorine bleaching lye,
i.e. sodium hypochlorite, NaOCl, solution. Details of the various biocides, the
production and storage of inorganic disinfectants, the chemistry and kinetics of
their use for seawater disinfection, and the design of the necessary dosing devices
can be found under [13] Chap. 2, Sect. 2.2.1 in Volume 2.

Chlorination
When dosing into the RO feed, primarily chlorine in the form of NaOCl solution is
used in this application and therefore only dosing of this biocide is discussed in detail
in the following. There is a distinction to be made depending on whether polyamide
spiral-wound or cellulose acetate hollow-fibre membrane elements are used in the
main desalination stage RO1 of the RO plant.
Chlorine has to be additionally injected into the RO feed if the residual chlorine
content resulting from chlorine dosing in the seawater extraction system and in the
pretreatment plant is not sufficient to prevent biological growth in the feed pipes and
the RO safety filtration equipment, or if chlorine has to be dosed into the feed line of
the RO membrane system, as is the case with cellulose membrane elements.
Chlorine dosing both to the SWRO plant and in the RO feed is usually intermit-
tent, for instance with dosing times of 10–60 min every 12–24 h operating time
whereby the manner of dosing, i.e. dosing sequence and Cl2 dosing rate, depends on
the seawater quality and the respective operating conditions (see [13] Chap. 2, Sect.
2.2.1 in Volume 2). During the dosing time, the chlorine concentration in the
pipeline to the safety filtration equipment should be at least 0.5 mg/l Cl2 to limit
biological growth within it. If this is not the case, chlorine will have to be subse-
quently dosed at an appropriate rate.
If the membrane system is equipped with polyamide spiral-wound elements, the
excess chlorine must be removed as far as possible by dechlorination to <0.1 mg/l
Cl2 to protect the membranes from oxidation. Bacteria and other microorganisms
682 5 Reverse Osmosis Membrane System: Core Process of SWRO

that have not been completely eliminated by the previous disinfection measures and
that reach the membranes lead to biological growth and biofouling, referred to as
“aftergrowth”. The extent of this depends on the amount of nutrients for the
biological fouling agents entering the membrane system. Investigations conducted
during trials and at operational membrane desalination plants have shown that it is
possible to reduce this nutrient supply by changing from continuous to intermittent
chlorination. The reason for this is that chlorination breaks down high-molecular
organic substances, which cannot normally be assimilated by bacteria, into smaller
molecules, thus making them available as nutrients at the membranes [28]. Due to
the shorter dosing time with intermittent dosing, the amount of assimilable nutrient
substance generated by chlorination is reduced accordingly.
If cellulose acetate membranes are used in the main desalination section, this
aftergrowth can be reduced by intermittently applying chlorine to the RO membrane
system for disinfection as these are more resistant to oxidation. When chlorine is
dosed in this way, which is designated by the cellulose hollow-fibre elements
manufacturer as intermittent chlorine injection (ICI), the membranes are fed inter-
mittently with chlorine at a dosing rate of 0.3–1.0 mg/l Cl2. The Cl2 dosing time τdos,
NaOCl depends on the biological activity of the feed as well as the desalination plant’s
operating conditions and ranges from 0.5 to 3 h per 24 h of RO operating time. The
residual chlorine content in the RO concentrate should not fall below 0.2 mg/l Cl2 to
ensure that the membrane elements will be swept by a sufficient excess of disinfec-
tant over their entire area. In this case, it is also possible to continuously chlorinate
the RO plant’s upstream equipment while providing intermittent chlorine dosing to
the membrane system by installing a dechlorination system just ahead of the RO
plant. This is then operated cyclically at the desired dosing rate, thus eliminating the
need for an additional chlorine dosing system at the RO feed.

Dosing Systems Sizing and Calculation of NaOCl Consumption The dosing


capacity FD, NaOCl that a chlorine dosing device needs to attain a defined chlorine
concentration cC l2,target is calculated from this target value, the flow rate FW of the
medium into which the chlorine is to be dosed, and the active chlorine content
cC l2,,NaOCl of the sodium hypochlorite solution (Eq. 5.244).

cC l2,target  F W
F D,NaOCl ¼ : ð5:244Þ
cC l2,NaOCl

FD, NaOCl ¼ dosing flow of NaOCl solution at delivery concentration [l/h]


cC l2,target ¼ target chlorine concentration for chlorination [mg/l ¼ g/m3]
cC l2,NaOCl ¼ available chlorine concentration of NaOCl solution [g/l]
When sodium chlorite solution is stored, its content of active chlorine decreases
with time (see [13] Chap. 2, Sect. 2.2.1 in Volume 2). When calculating the dosing
capacity with Eq. (5.244), the actual chlorine concentration FD, NaOCl at the time of
use must be taken into account as this is lower than the concentration at delivery
cC l2,,NaOCl .
5.5 RO Process Units, Components and Overall System Configuration 683

The maximum design dosing rate FD, NaOCl, X % , max of the chlorine dosing device
results from FD, NaOCl and the selected safety factor fsafety according to Eq. (5.244a).

F D,NaOCl, max ¼ F D,NaOCl  f safety : ð5:244aÞ

FD, NaOCl, max ¼ max. dosing flow of NaOCl solution at delivery concentration
[l/h]
If the sodium hypochlorite solution is to be dosed at a lower concentration than
the commercial concentration, the dilution water required for this is calculated using
Eq. (5.243) while the dosing rate of the diluted solution at the dosing point is
calculated together with the design values for the maximum dosing rate of the
diluted solution and the maximum dilution water flow required for this using
Eqs. (5.243a)–(5.243e).
The annual consumption of sodium hypochlorite solution CNaOCl,τo,y with inter-
mittent dosing is calculated as shown by Eqs. (5.244b) and (5.244c).

DC,NaOCl,τdos ¼ f dos,τ,NaOCl  F D,NaOCl : ð5:244bÞ


τdos,NaOCl
f dos,τ,NaOCl ¼ :
τo,NaOCl þ τdos,NaOCl

C NaOCl,τ0,y ¼ DC,NaOCl,τdos  τo,y : ð5:244cÞ

DC,NaOCl,τdos ¼ Amount of sodium chloride during dosing time [l/h]


fdos, τ, NaOCl ¼ dosing time factor for intermittent dosing of NaOCl
τdos, NaOCl ¼ NaOCl dosing time [min] [h]
τo, NaOCl ¼ operation period for NaOCl dosing time factor calculation [min] [h]
CNaOCl,τo,y ¼ consumption of sodium hypochlorite at delivery concentration
during annual operation time τo, y [l/year]
If the dosing time factor fdos, τ, NaOCl and the Cl2 dosing rate FD, NaOCl change
during the period of operation of the RO plant, the annual consumption of sodium
hypochlorite CNaOCl,τo,y is calculated as the sum of the quantity of sodium hypochlo-
rite consumed under the different dosing conditions as obtained using Eq. (5.241c).
The required storage capacity and the capacity of any holding tanks are calculated
according to Sect. 4.2.3.2, Eqs. (4.75)–(4.79).

Dechlorination
The RO feed can be dechlorinated using various reducing agents such as activated
carbon, hydrogen peroxide, thiosulphate, or sulphite compounds such as sodium
sulphite Na2SO3, sodium hydrogen sulphite/sodium bisulphite NaHSO3, or sodium
metabisulphite Na2S2O5. For RO desalination processes, sodium bisulphite and
sodium metabisulphite are used predominantly for dechlorination, so the chemistry
and the dimensioning of the dosing equipment are described in detail below only for
these compounds.
684 5 Reverse Osmosis Membrane System: Core Process of SWRO

The active compound for dechlorination is sodium bisulphite, NaHSO3. When


sodium metabisulphite Na2S2O5 is dissolved, it is converted to sodium bisulphite
NaHSO3 according to the reaction equation below.

S2 O5 2 þ H2 O ! 2HSO
3

Chlorine is present in solution as a mixture of hypochlorite ions OCl and


hypochlorous acid HOCl. Depending on pH, the proportion of the two compounds
shifts to OCl for a pH in the alkaline range and to HOCl for more acidic pH values
(see [13] Chap. 2, Sect. 2.2.1 in Volume 2). The reaction with bisulphite HSO3
differs accordingly.

OCl þ HSO 
3 ⇄SO4 þ Cl þ H
2 þ

HOCl þ HSO 
3 ⇄SO4 þ Cl þ 2H
2 þ

If the dechlorination reaction is referred to the sodium metabisulphite Na2S2O5


originally used, overall the following reaction equation results:

2OCl þ S2 O5 2 þ H2 O⇄2Cl þ 2SO2


4 þ 2H
þ

Due to the H+ ions released during the dechlorination reaction, the alkalinity of
the water is reduced during the reaction with bisulphite.
When seawater is chlorinated, the bromide Br present in a concentration of up to
70 mg/l is mostly oxidized to hypobromite OBr or hypobromic acid HOBr, but
partly also to elemental bromine Br2 while the dosed chlorine or hypochlorite is
reduced to inactive chloride (see [13] Chap. 2, Sect. 2.2.1 in Volume 2). In this case,
the disinfectant effect of the chlorine dosage is due largely to the bromine
compounds thus formed. During dechlorination of chlorinated seawater, the bromine
compounds therefore react for the most part with the added bisulphite. As far as the
reactions in total are concerned, however, the stoichiometric calculation can be
carried out with the equations shown above for the reaction of hypochlorite and
hypochlorous acid with bisulphite, as shown taking the following equations as
examples.

The reaction of bisulphite with the oxide compounds of bromine and chlorine
from a kinetic aspect is a fast reaction. However, dechlorination should be completed
within not more than 10–30 s and a residual halogen content of well below 0.1 mg/l
must be attained within this time. In order to achieve this, a sufficient surplus of
reducing agents is required.
5.5 RO Process Units, Components and Overall System Configuration 685

Dosing Systems Sizing and Calculation of Bisulphite Consumption Table 5.27


shows not only the stoichiometric factor fstoich, Reac resulting from the reaction
equations shown above and the stoichiometric quantity of reducing agent Ra per
gram of chlorine calculated from it, but also the excess factor fex, Ra as determined in
investigations and also used in practice for dimensioning the dosing devices for
dechlorination as well as, derived from this, the practical dosing quantity of reducing
agent Ra per gram of chlorine to be removed [141].
Accordingly, the required dosing rate RD, Ra, 100% for dechlorination is calculated
as the product of the stoichiometric factor fstoich, Reac and the excess factor fex, Ra, the
concentration of chlorine present cCl2 , and the ratio of the molecular weights of the
reducing agent MWRa and of chlorine MWCl2 (Eq. 5.245).

cCl2  MWRa
RD,Ra,100% ¼ f stoich,Reac  f ex,Ra  : ð5:245Þ
MWCl2
cCl2  MWRa
RD,Ra,100% ¼ f stoich,Reac  f ex,Ra  :
70:906
RD, Ra, 100% ¼ dosing rate of reducing agent at 100% concentration [mg/l ¼ g/m3]
fstoich, Reac ¼ stoichiometric factor of reducing agent []
cCl2 ¼ chlorine concentration to be reduced [mg/l ¼ g/m3]
MWRa ¼ molecular weight of reducing agent
MWCl2 ¼ molecular weight of chlorine
fex, Ra ¼ excess factor for reducing agent []
The dosing rate related to the delivery concentration of the reducing agent
solution is calculated using Eq. (5.245a). Sodium bisulphite and sodium
metabisulphite are obtainable in solid form and sodium bisulphite also as a solution
with a concentration of 35–38% and density of 1.3 kg/l.

RD,Ra,100%  100
RD,Ra,X% ¼ : ð5:245aÞ
X%del
RD,Ra,X% ¼ dosing rate of reducing agent at delivery concentration X%,del
[mg/l ¼ g/m3]
X%del ¼ delivery concentration of reducing agent [%]
However, when calculating the dosing rate, it is to be noted that the oxygen
contained in the water to be dechlorinated also consumes bisulphite.

O2 þ 2HSO
3 ⇄2SO4 þ 2H
2 þ

þ
O2 þ S2 O5 2 þ H2 O⇄2SO2
4 þ 2H
686
5

Table 5.27 Dechlorination—amounts of S(IV) compounds needed


Molecular
weight Stoichiometric factor Rate of Ra per g/l Cl2 Excess factor Rate of Ra per g/l Cl2 to
MW fstoich, Reac stoichiometric fex, Ra be applied
Reducing agent Ra Formula g/mol – g Ra/g Cl2 – g Ra/g Cl2
Sodium sulphite SS Na2SO3 126.044 1.0 1.778 ~3 ~5
Sodium bisulphite NaHSO3 104.062 1.468 ~2.0–2.3 ~2.9–3.4
SBS
Sodium Na2S2O5 190.109 0.5 1.341 ~2.7–3.1
metabisulphite SMBS
Reverse Osmosis Membrane System: Core Process of SWRO
5.5 RO Process Units, Components and Overall System Configuration 687

Table 5.28 Solution life Solution concentration Solution life


of bisulphite solutions
% NaHSO3 Months
[142]
10 0.25
20 1
30 6

The resulting additional amount of bisulphite needed is not included in the excess
factor fex, Ra of Eq. (5.245) and must be added to the dosing rate RD, Ra, 100%.
Oxidation of the bisulphite by oxygen also occurs during storage of the reducing
agent solution and likewise influences its sulphite content during prolonged storage,
whereby the degree of reduction of the bisulphite concentration depends on the
solution concentration of NaHSO3 (see Table 5.28). This should then be taken into
account when determining the dosing rate RD, Ra, X% by taking a correspondingly
lower value of the sulphite concentration X%del.
The nominal dosing rate of the reducing agent FD,Ra,X% results from Eq. (5.245b)
and the design maximum value of the dosing device rate FD, Ra, X % , max together
with the selected safety factor fsafety from Eq. (5.245c).

RD,Ra,X%  F W RD,Ra,100%100F W
F DRa,X% ¼ ¼ ð5:245bÞ
ρRaX%  1000 X%del  ρRaX%  1000

F D,RA,X%, max ¼ F DRa,X%  f safety : ð5:245cÞ

FD,Ra,X% ¼ dosing flow of reducing agent at delivery concentration of X%,del [l/h]


FD, Ra, X % , max ¼ max. dosing flow of reducing agent at delivery concentration of
X%,del [l/h]
ρRaX% ¼ density of bisulphite solution at delivery concentration of X%,del [kg/l]
If the supply or preparation concentration of the bisulphite solution is to be
diluted to a lower dosing concentration, the required dilution water rate and also
the resulting nominal and maximum dosing rate can be calculated with
Eqs. (5.243a)–5.243e).
When calculating the annual consumption of bisulphite C C,Ra,X%,τo,y , the intermit-
tent dosing mode of the reducing agent depending on the chlorination dosing
frequency must be taken into account. Dosing of bisulphite starts slightly earlier
than the start of chlorination, so the bisulphite dosing time is longer than that of
chlorine. When the dosing time factor fdos, τ, Ra for bisulphite dosing is taken into
consideration, its annual consumption is calculated with Eqs. (5.245d) and (5.245e.

RD,Ra,X%  F W
DC,Ra,X%deliv,τdos ¼ f dos,τ,Ra  : ð5:245dÞ
1000
τdos,Ra
f dos,τ,Ra ¼ :
τo,Ra

C C,Ra,X%,τo,y ¼ DC,Ra,X%deliv:,τdos  τo,y : ð5:245eÞ

DC,Ra,X%del,τdos ¼ amount of reducing agent at delivery concentration of X%,del


during dosing time [kg/h]
688 5 Reverse Osmosis Membrane System: Core Process of SWRO

fdos, τ, Ra ¼ dosing time factor for intermittent dosing of reducing agent


τdos, Ra ¼ dosing time of reducing agent [min] [h]
τo, Ra ¼ operation period for reducing agent dosing time factor calculation [min]
[h]
τo, y ¼ annual operation time of RO system [h/year]
CC,Ra,X%del,τo,y ¼ consumption of reducing agent at delivery concentration of X%,
del during annual operation time τ o, y [kg/year]

If the dosing rate and dosing time factor for dosing of the reducing agent change
during the annual RO operating time, then the annual consumption of bisulphite is
calculated as the sum of the consumptions during each operating period for a
constant dosing cycle (see Eq. 5.241c).
Details of the calculation of the capacities of the sodium bisulphite storage tanks
and the dosing tanks, should these be needed, can be found under Sect. 4.2.3.2,
Eqs. (4.75)–(4.79).

Location of Chlorination and Dechlorination Dosing Points


If additional dosing of chlorine into the feed section of an RO plant is required, the
dosing point should be located upstream of the safety filters to ensure that a
sufficiently high excess of disinfectant will be available for these. This prevents
solids accumulating in the filters from becoming biologically active and giving rise
to biological fouling right at the filtration stage. Any “aftergrowth” already occurring
in the safety filters reduces their service life and can also have the consequence that
more microbes from the filters will enter the RO membrane system. This means that
for dechlorination, the bisulphite dosing point should also be located downstream of
the safety filters. This is also accepted practice for RO systems.
However, investigations at operating plants also show that with this configuration
of the dechlorination point, although the safety filters have a longer service life than
if dechlorination takes place before the filters, there is increased aftergrowth in the
membrane system [143]. Based on these findings, the membrane manufacturer
Hydranautics recommends that dechlorination in the RO feed should be located as
far away as possible from the membrane system [144]. It is evident that the optimal
configuration for dechlorination is strongly dependent on the seawater’s biological
activity and composition as well as the desalination plant’s operating conditions. It is
therefore recommended, as shown in Fig. 5.63, to provide locations for dosing
bisulphite both upstream and downstream of the safety filters so as to have optimi-
zation options for minimizing biofouling during plant operation.
If the main desalination stage RO1 of an RO tract is equipped with cellulose
acetate membranes but the post-desalination stage RO2 is equipped with polyamide
membranes and in RO1 the CA membranes are intermittently swept with chlorine as
described above, the RO1 product water must be dechlorinated before piping it to
RO2. The dechlorination dosing point will then have to be located upstream of the
point in the RO2 feed line where the pH is raised.
5.5 RO Process Units, Components and Overall System Configuration 689

5.5.2.4.3 Power Demand of RO Systems Chemicals Dosing


The total power demand PDCi, X % , dos, T of the chemical dosing equipment for
conditioning the feed to the main desalination section RO1 and for post-desalination
RO2 is made up of the shares for continuous dosing of antiscalant and acid in the
RO1 feed and of sodium hydroxide in the RO2 feed as well as intermittent dosing of
hypochlorite in RO1 and, if necessary, bisulphite in the RO2 feed (Eq. 5.246).

X
i¼n X
i¼n
PDCi,X%,dos,T ¼ PDCi,X%,dos,cont þ PDCi,X%,dos,interm : ð5:246Þ
i¼1 i¼1

PDCi, X % , dos, T ¼ total power demand of dosing systems for chemical i in the
RO1 and RO2 feeds [kW]
PDCi, X % , dos, cont ¼ power demand for continuous dosing of chemical i [kW]
PDCi, X % , dos, interm ¼ power demand for intermittent dosing of chemical i [kW]
For the continuous dosing systems, the power takeup results from the dosing rate
FDCi, X % , dos at the dosing concentration X%,dos at the dosing point for this chemical
and the prevailing operating pressure there pCi, d, p as well as the overall efficiency
ηPDi,T of the respective dosing pump in accordance with Eq. (5.246a). The power
demand for intermittent dosing PDCi, X % , dos, interm is calculated in a similar way,
although additionally by multiplying with the respective dosing time factor fdos, τ, i,
as shown by Eqs. (5.245d) and (5.246b).

F DCi,X%,dos ,  pCi,d,p
PDCi,X%,dos,cont ¼ : ð5:246aÞ
36  ηPDi,T

F DCi,X%,dos ,  pCi,d,p
PDCi,X%,dos,interm ¼ f dos,τ,i  : ð5:246bÞ
36  ηPDi,T

FDCi, X % , dos ¼ dosing flow of chemical i at dosing concentration X%dos [m3/h]


pCi, d, p ¼ pressure at dosing point of chemical i [bar]
ηPDi,T ¼ total efficiency of dosing pump of dosing station for chemical i []
fdos, τ, i ¼ dosing time factor for intermittent dosing of chemical i []

5.5.2.4.4 Control of Rate, Flow, and Sequence of Chemicals Dosing


The addition of chemicals both into the feed line to the main desalination section
RO1 and to that of post-desalination RO2 must be primarily proportional to the
respective flow rate. In addition, the dosing rate must be adapted to the composition
of the feed water, the specified target values for dosing of chemicals, and the
operating conditions of the membrane systems. If positive displacement pumps
such as reciprocating and diaphragm metering pumps are used, the dosing rate can
be adjusted for these two controlled variables in proportion to the quantity to be
dosed by changing the speed of the metering pumps, while the dosing amount, which
is based on the target dosing value, is adjusted by changing the piston stroke.
690 5 Reverse Osmosis Membrane System: Core Process of SWRO

With antiscalant dosing to the feed lines of RO1 and RO2, the dosing rate is
matched to the flow rate by modulating the dosing pump speed, while the antiscalant
dosing rate is preset manually via the piston stroke and, if necessary, adjusted
accordingly.
If acid is dosed to the RO1 feed, in addition to the flow-proportional control, the
actual pH is measured by an online pH probe and the amount of dosed acid is adjusted
by changing the piston stroke to attain the target pH. The pH value in the RO2 feed is
regulated similarly by sodium hydroxide dosing.
If bisulphite is dosed additionally, its flow-proportional control is likewise
affected by varying the dosing pumps’ speed. Dechlorination, though, is oriented
to the dosing sequence specified for chlorination. Dosing of the reducing agent
commences in advance of the start of chlorine dosing. The amount of bisulphite
dosed is preset by the piston stroke and the degree of dechlorination, or excess
sulphite, is measured by an online probe for determining the oxidation-reduction
potential (ORP). The measured ORP value should be below 175–200 mV to achieve
a residual chlorine content of less than 0.1 mg/l Cl2. To meet this target, the piston
stroke of the bisulphite dosing pump and thus the NaHSO3 dosing rate is adjusted by
the control system.

5.5.2.5 Membrane Flushing, Cleaning, and Preservation


Even with effective pretreatment upstream of the RO system, chemical conditioning
at the RO feed to prevent scaling, and compliance with the maximum silt density
index SDI or modified fouling index MFI as specified by the membrane
manufacturers, the membranes will still become clogged during the desalination
plant’s operating time. The consequence of this is that there will not only be an
irreversible reduction in product performance and salt retention in the plant due to
membrane ageing, but also a reversible reduction due to this membrane clogging.
This reversible reduction in performance can be undone again to a large extent by
membrane cleaning (see Sect. 5.3.3.2.2, Fig. 5.49). For such membrane treatment,
the plant’s RO unit that is to be cleaned is taken out of operation and swept with
cleaning solutions over a lengthy period via a cleaning station for clean-in-place
(CIP) membrane cleaning. In line with membrane manufacturers’ specifications,
such membrane treatment must be done at the latest when the normalized values of
the plant’s product flow and salt passage as well as the membrane modules’
differential pressure deteriorate by a specified value due to fouling. Table 5.29
shows the respective ranges for these parameters as specified by various membrane
suppliers as an indicator for commencing cleaning. The membranes must be cleaned
as soon as any one of these limit values is attained.
However, even if these criteria for the start-up of membrane cleaning are com-
plied with, during a lengthy period of operation a fouling deposit has already formed
on the membranes. Through predictive or preventive cleaning for which cleaning
takes place more frequently and already below the limit values of the operating
parameters specified by the membrane manufacturers, it is possible to prevent the
residues on the membranes from excessive aging and thus resulting in the need to
increase the cleaning effort in terms of duration and intensity [145]. A further
5.5 RO Process Units, Components and Overall System Configuration 691

Table 5.29 Chemical cleaning—membrane manufacturer’s cleaning start-up criteria


Cleaning start–up
Operation parameter Unit criterion Note
Normalized product flow m3/day; Loss of 10–15% Of initial capacity or after last
m3/h cleaning
Normalized differential bar Increase of 10–20%
pressure
Normalized salt passage % Increase of 5–10%

advantage of this cleaning method is on average lower pressure loss of the membrane
modules and higher product performance, i.e. also lower energy consumption of the
plant and also better permeate quality.
The residues that accumulate on the surfaces of the membrane system during
reverse osmosis desalination are a mixture of organic and inorganic components.
With surface seawater extraction in particular, these are to a larger extent organic
substances that can be biologically active or inactive, i.e. organic and biological
foulants. In addition, there are inorganic components such as heavy metal
hydroxides, in particular iron oxides and iron hydroxides from the dosage of iron
(III) salts as flocculants in the flocculation stage of pretreatment, as well as inorganic
solids naturally occurring in seawater, like silt and clay, that pass through
pretreatment in colloidal form. If the dosing rates of acid or antiscalant are insuffi-
cient, especially if dosing devices malfunction, scalants such as alkaline earth
sulphates and calcium carbonate may also be part of the inorganic make-up of the
coatings. Channel formation in the spacers of the membrane elements caused by
fouling with locally intensified concentration polarization can also result in local
scaling [146].

5.5.2.5.1 Membrane Flushing and Cleaning Methods


The parameters of product flow, differential pressure, and salt passage listed in
Table 5.29 can be improved to some extent by flushing the membrane system with
feedwater or permeate at low pressure. With feedwater flushing, the RO operating
pressure is lowered by opening the concentrate valve of the membrane system as far
as the feed pumps allow. Permeate flushing is done by running the cleaning/flushing
pumps of the RO system cleaning station (Fig. 5.99) at the maximum flushing
capacity permitted for the membrane modules (Table 5.32) at an operating pressure
of not more than 4–5 bar.
However, in order to remove reversible fouling from the membranes to a great
extent, their treatment with chemical cleaning solutions is necessary. Table 5.30
summarizes the basic formulations of the cleaning solutions proposed by membrane
manufacturers to remove particular types of fouling.
A distinction has to be made between polyamide membranes and membranes of
cellulose acetate with regard to the allowable pH range in chemical cleaning. With
the former, cleaning is possible within a pH range of 1–12, while with CA/CTA
membranes the pH range when cleaning is limited to 3–8. However, also for
692 5 Reverse Osmosis Membrane System: Core Process of SWRO

Permeate Chemicals
Concentrate
from RO

Permeate to drain or
from RO treatment
Cleaning
solution
to RO

PH dP

TIC Heat
LIC exchanger PI FI

Cleaning
tank
Cleaning /
flushing
pump Cartridge filter
Permeate
to drain or
treatment

Fig. 5.99 Chemical cleaning—cleaning station

polyamide membranes in the strongly acidic and strongly alkaline ranges, the
permissible pH value is sometimes restricted to varying extents by the membrane
manufacturers depending on the temperature of the cleaning solution (Table 5.31). In
some cases, with very high acidity and alkalinity, the permissible cleaning time is
also limited at elevated temperatures.
In addition to pH, though, the actual composition of the cleaning solution also
influences the maximum permissible temperature and pH during cleaning as
suggested by the membrane manufacturer, so for this reason the proposed operating
values for these parameters are likewise stated in Table 5.30 for the formulation of
the cleaning solution.
Cleaning performance generally is enhanced at higher temperature and thus RO
cleaning stations should be equipped with a heating system to adjust the temperature
of the cleaning solution.
Organic and biological coatings on the membranes can be removed most effec-
tively using highly alkaline cleaning solutions. Surfactants like detergents are added
to these solutions to improve the removal of residues from the membrane surface and
their dispersion in the cleaning solution. This product group of organic anion-active
surfactants includes the compounds sodium dodecyl sulphate (SDS), sodium
dodecyl benzene sulphate (Na-DDBS), and polyoxyethylene sodium laurylsulphate
(PSLS). But also inorganic compounds such as sodium tripolyphosphate (STPP)
have dispersing properties.
Acid-soluble compounds such as the alkaline earth carbonates CaCO3, CaMg
(CO3)2, and MgCO3 can be removed by acidic cleaning solutions such as
hydrochloric acid, phosphoric acid, or citric acid in a pH range from 1 to 4. The
alkaline earth sulphates CaSO4, SrSO4, and BaSO4 are only slightly or not at all acid
5.5 RO Process Units, Components and Overall System Configuration 693

Table 5.30 Chemical cleaning—membrane contaminants and recommended cleaning solutions


Cleaning conditions
Cleaning solution Temperature
Contamination composition pH ( C) Notes
Polyamide membranes
Alkaline earth Hydrochloric acid 0.2– 1–2.5 25–35
carbonates 2%
CaCO3, CaMg Citric acid 1–2% 2–4 40 pH adjustment
(CO3)2, MgCO3 (+NH4OH) also with NaOH,
also without pH
adjustment
Sodium dithionite 4–6 25
(sodium hydrosulphite)
Na2S2O4 1%
Phosphoric acid H3PO4 1–2 25
0.5%
Alkaline earth Sodium tripolyphosphate 10–12 30–40 pH adjustment
sulphates (STPP) 2% + Na-EDTA also with sodium
(CaSO4, SrSO4, 0.8–1% phosphate or
BaSO4) NaOH 0.1%
Sodium 2 35
hexametaphosphate
SHMP 1% + HCL
Hydrochloric acid HCl 2.5 35
2%
Metal oxides/ Citric acid 2% 2 40
hydroxides (Fe, Sodium dithionite 5 30
Mn, Cu etc.) (sodium hydrosulphite)
Na2S2O4 1%
Phosphoric acid H3PO4 1–2 25
0.5%
Sulfamic acid NH2SO3H 3–4 25
1%
Mixed Sodium tripolyphosphate 10–12 30–40
inorganic/ (STPP) 2% + Na-EDTA
organic colloids 0.8–1%
NaOH 0.1% + sodium 11–12 30–35
dodecyl sulphate (SDS)
0.03–0.2%
Hydrochloric acid 0.2% 2.0–2.5 35–40 Following the
alkaline cleaning
Biological and Sodium hydroxide 11.5–13 30–35
organic matter NaOH 0.1%
Sodium tripolyphosphate 10–12 30–40
(STPP) 2% + Na-EDTA
0.8–1%
Sodium tripolyphosphate 10 35–40 pH adjustment
(STPP) 2% + sodium also with NaOH
0.1%
(continued)
694 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.30 (continued)


Cleaning conditions
Cleaning solution Temperature
Contamination composition pH ( C) Notes
dodecyl benzene sulphate
(Na-DDBS) 0.025%
NaOH 0.1% + sodium 11–12 30–35
dodecyl sulphate (SDS)
0.03–0.2%
NaOH 9–11 30–35
0.1% + polyoxyethylene
sodium laurylsulphate
(PSLS) 0.1–0.5%
Hydrochloric acid 0.2% 2.0–2.5 35–40 Following the
alkaline cleaning
Cellulose acetate membranes
Calcium Citric acid 2% + NH4OH 4.0 30–40
carbonate, metal
oxides/
hydroxides (Fe,
Mn, Cu etc.),
inorganic
colloids
Biological Chlorine 5–10 mg/l Cl2 5–6 Intermittent
matter/organics chlorination with
increased Cl2
concentration in
feed or shock CIP
cleaning
Citric acid 2% + NH4OH 4.0
Ultrasil 53 1%

Table 5.31 Chemical cleaning—temperature and pH limits during cleaning


Temperature pH range limits during cleaning
SW membranes BW membranes

C Manufacturer A Manufacturer B Manufacturer A Manufacturer B
25 1–13 1–12 1–13
35 1–12 1–11 1–12
45 2–11 1–10.5 2–10.5 1–10.5

soluble and are detached from the membrane surface by strong chelating agents such
as Na-EDTA (tetra-sodium salt of ethylene diamine tetraacetic acid). Citric acid has
both chelating and acid dissolving properties. This compound is therefore used to
remove deposits of metal oxides and hydroxides together with alkaline earth
carbonates.
In practice, membrane coatings are made up of a mixture of different types of
contamination. The components of the fouling deposits may vary depending on
5.5 RO Process Units, Components and Overall System Configuration 695

seasonal fluctuations in seawater quality and also on how effective pretreatment has
been. It is therefore necessary that:

• either a combination of formulations is used that can effectively remove both


organic and inorganic fouling
• or the fouling layers be treated successively with several contamination-specific
chemicals.

Which fouling products are deposited can be determined by conducting an


autopsy of the fouled membrane elements and analysing the residues. Another
possibility is to perform trials with fouled elements on a pilot scale with different
cleaning solutions, analyse the solutions after cleaning is completed, and check the
respective cleaning effect using a standard NaCl solution in a pilot plant (see Sects.
5.1.5.2.3 and 5.3.3.2.2). Such a trials rig should therefore be part of the lab
equipment of an SWRO facility.
Fouling-specific membrane cleaning agents made up of many active components
are offered by specialist companies that also manufacture antiscalants and
flocculants for RO plants. An example of such an agent is Ultrasil 53, which the
membrane manufacturer Toyobo recommends for removing organic and biological
fouling from its CTA membranes. In addition to surfactants and chelating agents,
this also contains enzymes to break down organic substances during cleaning. When
using membrane cleaning agents from specialist companies, the compatibility of the
selected product with the membranes used in the plant should be confirmed by the
membrane manufacturer.

5.5.2.5.2 Chemical Cleaning Station Design


A cleaning station (Fig. 5.99) consists of:

• the cleaning tank in which the chemical solutions are made up


• the cleaning and flushing pump
• a heat exchanger either integrated into the cleaning tank or installed at the feed
line for setting the temperature of the circulating cleaning solution
• a 5–10 μm cartridge filter in the feed pipe to the membrane system
• the supply and return lines for the cleaning solution
• instrumentation and control equipment for regulating:
– pH and temperature in the cleaning tank and of the circulating cleaning
solution
– the liquid level in the cleaning tank
– the flow rate and pressure in the feed line
– the differential pressure across the cartridge filter.

The cleaning solution is made up in the cleaning tank using permeate either from
the main desalination stage or, in the case of a two-pass system, from post-
desalination.
696 5 Reverse Osmosis Membrane System: Core Process of SWRO

The volume of the solution required for cleaning the membranes Vcleaning is
calculated from the volume of cleaning solution required for a single membrane
Vsolut.element, and the number of membrane elements to be cleaned NE,cl as well as the
volume of the cartridge filter, the volume of the feed line to the membrane system,
and the volume of the return line Vpiping + CF (Eq. 5.247). The number of membrane
elements to be cleaned NE,cl also results from the number of membrane modules to
be treated NM,cl and the number of membrane elements with which they are equipped
NE,M (Eq. 5.247a). The volume of cleaning solution Vcleaning so calculated is also
equal to the required net capacity of the cleaning tank VCt,net.

V cleaning ¼ V Ct,net ¼ V solut,element  N E,cl þ V pipingþCF : ð5:247Þ

N E,cl ¼ N M,cl  N E,M : ð5:247aÞ

Vcleaning ¼ cleaning solution volume [l] [m3]


VCt,net ¼ net volume of cleaning tank [l] [m3]
Vsolut.,element ¼ cleaning solution requirement per element [l] [m3]
NE,cl ¼ number of elements to be cleaned []
Vpiping + CF ¼ volume of cleaning feed and cartridge filter and return piping
[l] [m3]
NM,cl ¼ number of modules/vessels to be cleaned []
NE,M ¼ number of elements per module/vessel []
In an SWRO plant in a two-pass configuration, the cleaning station is dimen-
sioned on the basis of the main desalination stage, as the number of membrane
elements and membrane modules installed there is significantly higher than for post-
desalination.
The membrane manufacturers base their calculations of the volume of the
cleaning solution Vcleaning on differing values for the required element-specific
volume. The membrane manufacturer DuPont Filmtec calculates Vcleaning based on
the empty volume of the membrane elements Vempt,element according to the values
listed in Table 5.32, column 1 and the resulting empty volume of the membrane
modules. Other membrane manufacturers state higher values for the required solu-
tion volume per element Vsolut.element, whereby these are graded in part depending on
the degree of fouling (Toray) (Table 5.32, Column 2).
The membrane manufacturer Toyobo quotes the following values for its CTA
hollow-fibre membrane module HU10255 according to Table 5.33 for the solution
volume per element, the flow rate during flushing and cleaning, and the pressure
conditions to be maintained.
Right at the start of the cleaning process, the water in the feed and return lines of
the cleaning station as well as in the membrane system itself must be displaced with
cleaning solution. A residual amount of cleaning solution remains in the cleaning
tank, this being what is termed the usable volume for cleaning VCt,use. If dilution of
the cleaning solution is to be avoided, the required filling volume Vfilling of the RO
unit to be cleaned is the sum of the pipeline volumes and the volume of the cartridge
filter Vpiping + CF of the cleaning station plus the water volume of the membrane
5.5 RO Process Units, Components and Overall System Configuration 697

Table 5.32 Chemical cleaning—spiral-wound membrane element volumes and cleaning/flushing


rates per vessel
Cleaning
Empty and
membrane Cleaning solution Water volume flushing
element vessel volume of membrane flow rate
Element diameter and volume requirement element vessel Fflush,
length Vempt,element Vsolut.,element Vwater,element module
m3/h,
l/element l/element l/element module
inches mm 1 2 3 4
4  40 100a  1016 8 10–15 5 1.5–3.0
8  40 201a  1016 33 40–60 20 6.0–12.0
16  402a  1016 132 136–240 80 33.0–44.0
40
a
Internal element diameter

Table 5.33 Chemical cleaning—CTA hollow-fibre membrane element volumes and cleaning/
flushing rates per module
Cleaning and
flushing flow rate,
min. Cleaning and Cleaning solution volume
Fflush/clean,module flushing pressure requirement Vsolut.,element
l/min, m3/h,
Procedure module module bar l/module
Flushing 30 1.8 min. 2 –
Chemical 52 3.1 min. 3, max. 5 300
cleaning

elements to be cleaned Vwater,element as shown in Table 5.32, Column 3 (see


Eqs. 5.247b and 5.247c). If there is to be no dilution of the cleaning solution, the
filling factor ffilling is unity. If partial dilution is acceptable, the value of ffilling is
reduced accordingly, resulting in a lower filling volume Vfilling.

V Ct,use ¼ V cleaning  V filling : ð5:247bÞ


 
V filling ¼ f filling  V pipingþCF þ V Water,element 3 N E,cl : ð5:247cÞ

VCt,use ¼ usable volume in cleaning tank [l] [m3]


Vfilling ¼ filling volume of membrane elements and cleaning piping [l] [m3]
ffilling ¼ filling factor ¼ 1 for no dilution of cleaning solution []
Vwater,element ¼ water volume of element [l] [m3]
After the cleaning and membrane systems have been filled with cleaning solution,
the residual level of the solution in the cleaning tank must be sufficiently high so that
698 5 Reverse Osmosis Membrane System: Core Process of SWRO

the available NPSHa for the flushing and cleaning pump is greater than the NPSHr
value specified by the manufacturer of the cleaning and flushing pumps (see Sect.
5.5.2.2.2). If this is not the case, either the volume of the cleaning solution Vcleaning
must be increased, i.e. the value of the cleaning volume Vsolut.element specific to the
membrane elements must be increased, or else the filling volume Vfilling, i.e. the
filling factor ffilling, must be reduced. If the volume of the cleaning solution Vcleaning is
increased, the net volume VCt,net of the cleaning container will also increase. If the
filling factor ffilling is reduced, the cleaning solution is diluted in the systems of the
RO unit accordingly, unless a more concentrated solution is made up or during
recirculation the concentration specified for cleaning is restored by adding
chemicals.
The delivery Fflush,max of the flushing and cleaning pump is calculated on the
basis of the permissible flow rate Fflush,module specified by the membrane
manufacturers for the membrane modules according to Table 5.32, Column 4 and
the number of membrane modules NM,cl to be cleaned (Eq. 5.248).

F flush, max ¼ F flush,module  N M,cl : ð5:248Þ

Fflush,max ¼ max. cleaning/flushing design flow of cleaning pump [m3/h]


Fflush,module ¼ max. cleaning/flushing flow per module/vessel [m3/h]
On the basis of the resulting delivery Fflush,max of the flushing and cleaning pump,
a check must then be made as to whether a retention time τret of 3–5 min is achieved
during recirculation with the usable residual volume of cleaning solution VCt,use in
the tank (Eq. 5.249).

F flush, max  τret


V Ct,use ¼ : ð5:249Þ
60
τret ¼ retention time of cleaning tank useable volume ¼ 3–5 [min.]
Permeate flushing and circulation of the cleaning solution are done at low
pressures, i.e. for the polyamide spiral winding elements within a pressure range
of 1–4 bar, for CTA hollow-fibre modules according to the values in Table 5.33. The
cleaning and flushing pumps must be dimensioned accordingly.
For chemical cleaning of a membrane unit with 199 membrane modules, each
equipped with eight 800 elements, i.e. 1592 elements for cleaning in total, and with a
cleaning volume per element of 40 l/element, a net volume of the cleaning tank of
approx. 160 m3 is required according to Eqs. (5.247) and (5.247a, taking into
account the volumes of the cleaning station cartridge filter and the pipes connecting
the cleaning station to the unit to be cleaned. The cleaning and flushing pumps must
be designed for a maximum flushing delivery of approx. 2000 m3/h for a maximum
flushing flow rate of 10 m3/membrane module.
The design of the cleaning station may show that the capacity and dimensions of
the cleaning tank when designed to cater for the main desalination stage RO1 results
in a usable volume VCt,use that is insufficient, i.e. a filling level that is too low for
cleaning the post-desalination stage RO2, which is equipped with fewer membrane
5.5 RO Process Units, Components and Overall System Configuration 699

modules. This will be the case in particular if the RO2 post-desalination unit with
more than one stage is not treated in a single cleaning procedure but stage wise one
stage after the other. If a multi-stage membrane system is severely fouled and all its
stages were to be cleaned in a single operation, the contamination of the first more
heavily fouled stage may be transported to subsequent stages, and for this reason
treatment is done in steps. If cleaning is done in this way, the quantity of cleaning
solution and the cleaning pump delivery must be based on the number of modules of
the particular stage to be cleaned. So that the cleaning station can then be used for the
entire RO tract, the size of the segments to be cleaned in the main desalination stage
RO1 must be reduced accordingly, which means that the array to be treated there
must be broken down into sections that are cleaned one after the other. Although this
reduces the cleaning station’s tank capacity and pump delivery, the time required for
cleaning a unit and thus also the downtime of this part of the RO tract is prolonged
depending on how many sections there are. This also applies to the downtime for
step-by-step cleaning.
However, such severe contamination of the post-desalination stage RO2 that
would necessitate step-by-step cleaning there is normally hardly encountered in
seawater desalination systems except that strong scaling is formed in the post-
desalination due to alkaline operation.

5.5.2.5.3 Cleaning Procedure and Cleaning Sequences


The sequence of a cleaning procedure is shown in Table 5.34.
At the start of the cleaning process, it is recommended to flush the membrane
system with pretreated seawater at low pressure via the feed pipe to remove loose
deposits (Step 2). Then the membranes are flushed with permeate to displace all
seawater and concentrate from the system (Step 3). This is particularly important
when cleaning is to be carried out with an alkaline solution so as to prevent
precipitation that is formed from residual seawater and concentrate during the
cleaning procedure. At the same time as these flushing steps, the cleaning solution
is prepared in the cleaning tank and heated to the temperature specified for the
cleaning process (Step 1). Permeate flushing is followed by filling the membrane
system with the cleaning solution and displacing the flushing water so that the
solution is diluted as little as possible (Step 4). The cleaning solution from the
cleaning tank is then circulated through the membrane system and the temperature in
the circuit is regulated via the heat exchanger in the cleaning station (Step 5). If
necessary, chemicals can be added in this step and the solution’s pH can also be
monitored. Depending on the extent of membrane fouling, the delivery rate and
pressure of the cleaning pump are adjusted so that the pressure range specified for
cleaning is not exceeded. After a specified time period, circulation is stopped. The
solution now remains stagnant in the system in order to seep into and soften the
residues on the membranes for a defined period, referred to as the soaking time (Step
6). This soaking time is followed by renewed circulation, this time at the full output
of the cleaning/flushing pump and with readjustment of the temperature in the circuit
(Step 7). Then the cleaning solution is drained from the membrane system and the
cleaning tank and the tank is rinsed with permeate (Step 8). Permeate is then used to
700 5 Reverse Osmosis Membrane System: Core Process of SWRO

Table 5.34 Chemical cleaning procedure—sequence and duration of cleaning steps


Step Cleaning sequence Duration for one RO stage
1 Cleaning solution preparation and heating –
(if necessary)
2 Low-pressure feed water flushing τFF1 ~15 min
3 Permeate flushing and displacement of feed/ ~15 min
brine τPF1
4 Cleaning solution filling and membrane system ~20 min
water content displacement τfill
5 Cleaning solution circulation τcirc1 ~30 min
6 Soaking ~30–60 min. up to some hours
(depending on degree of fouling)
7 Cleaning solution circulation τcirc2 ~30 min
8 Cleaning tank and membrane system drain ~30 min
9 Permeate flushing of membrane system and ~30 min
rinse of cleaning solution τPF2
10 Restart of membrane system and service rinse ~20–60 min up to some hours
with feed water τFF2 depending on type of cleaning
11 Switching back to normal RO operation mode ~20 min.
Minimum operational outage Up to step 9
200–230 min 3–4 h
Up to step 11
240–310 min 4–5 h

flush the residual cleaning solution out of the membrane modules and the cleaning
circuit (Step 9). Although the actual cleaning procedure is completed with this step,
feed water is again admitted via the feed pipe to the membrane system, which is then
brought up to operating pressure while the resulting permeate is discharged until the
specified conductivity of the membrane unit is again attained (Step 10). The
membrane system is then returned to normal operation (Step 11).
The sequence of cleaning steps with their respective approximate durations is
shown in Table 5.34. It should be borne in mind, though, that cleaning procedures
are highly dependent on the amount and type of fouling to be removed and the extent
to which these residues have aged. These times can therefore only be very rough
indicative values and they should merely serve as a guide for the approximate time
frame for cleaning an RO unit. In particular, the time required for soaking may range
from about 30 min for minor fouling to several hours for membranes with heavy
organic/biological fouling. In the most straightforward case, the time taken for the
cleaning procedure until the membrane unit can be returned to operation can be
around 3–4 h. This, though, refers only to the unit’s downtime and does not include
the time needed for preparations and making up the cleaning solutions.
However, the coatings to be removed consist of a mixture of inorganic and
organic/biological components. Therefore, often multiple cleaning procedures car-
ried out in succession will be necessary. Even with alkaline cleaning solutions as
needed to remove organic and biological fouling, the membranes have to be treated
5.5 RO Process Units, Components and Overall System Configuration 701

with an acid solution to shorten the service rinse time. If two cleaning sequences are
required within one cleaning procedure, Steps 1–9 of Table 5.34 are carried out in
the first cleaning sequence followed by repeating Step 1 and Steps 3–11 in the
second sequence. Cleaning will then take 7–9 h. If arrays in the main desalination
stage RO1 are broken down into sections, the cleaning time increases with their
number. For cleaning in steps, the cleaning time increases with the number of steps.

5.5.2.5.4 Water, Chemicals, and Energy Consumption During Cleaning


Seawater and Permeate Consumption
Pretreated seawater is consumed during low-pressure flushing before actual mem-
brane cleaning begins (Step 2) and at the end of the cleaning procedure during
service flushing (Step 10). The water volume required for this is calculated using
Eq. (5.250).

1
V SW,clp ¼  ðF FF2,i  τFF2,i þ F FF10,i  τFF10,i Þ: ð5:250Þ
60
VSW,clp ¼ Seawater volume consumption during cleaning procedure [m3]
FFF2, i ¼ Flushing flow during low-pressure flushing Step 2 [m3/h]
τFF2, i ¼ Duration of low-pressure flushing Step 2 [min]
FFF10, i ¼ Flushing flow during low-pressure flushing Step 10 [m3/h]
τFF10, i ¼ Duration of service rinse Step 10 [min]
Permeate is used in membrane cleaning for making up the cleaning solution (Step
1), for permeate rinsing to displace seawater and concentrate from the membrane
modules (Step 3), and for flushing the membrane system in Step 9 (see Table 5.34).
The permeate volume VP,i consumed per cleaning sequence is calculated accordingly
with Eqs. (5.251 and (5.251a.

V P,i ¼ V P3,i þ V P9,i þ V cleaning,i : ð5:251Þ

1
V P,i ¼  ðF PF3,i  τPF3,i þ F PF9,i  τPF9,i Þ þ V cleaning,i : ð5:251aÞ
60
VP,i ¼ Permeate volume consumption during cleaning sequence i [m3]
FPF3, i ¼ Flushing flow during permeate flushing Step 3 [m3/h]
τPF3, i ¼ Duration of permeate flushing Step 3 [min]
FPF9, i ¼ Flushing flow during permeate flushing Step 9 [m3/h]
τPF9, i ¼ Duration of permeate flushing Step 9 [min]
Vcleaning, i ¼ Cleaning solution volume for cleaning sequence i [m3]
If the cleaning procedure consists of a number ncl,s of sequences, the total
permeate volume per cleaning procedure VP,clp is calculated from the sum of the
volumes of permeate VP,i consumed in each sequence (Eq. 5.251b).
702 5 Reverse Osmosis Membrane System: Core Process of SWRO

X
i¼ncl,s

V P,clp ¼ V P,i : ð5:251bÞ


i¼1

VP,clp ¼ total volume of permeate consumed during the cleaning procedure [m3]
ncl,s ¼ number of cleaning sequences during the cleaning procedure []
The annual consumptions of seawater VSW,y and permeate VP,y for cleaning the
RO system are then calculated from their respective volumes per cleaning procedure,
VSW,clp and VP,clp, and the number of cleaning procedures per year nclp,y (Eqs. 5.252
and 5.252a).

i¼n
X clp,y

V SW,y ¼ V SW,clp : ð5:252Þ


i¼1

i¼n
X clp,y

V P,y ¼ V P,clp : ð5:252aÞ


i¼1

VSW,y ¼ annual seawater volume consumption during cleaning [m3/year]


VP,y ¼ annual permeate volume consumption during cleaning [m3/year]
The number of cleaning procedures per year nclp,y in turn is calculated from the
annual number of cleaning procedures for each unit to be cleaned nclp, y, unit and the
number of units Nunits f. cleaning of the RO tract to be cleaned during an operating year
(Eq. 5.252b). With effective pretreatment of the RO feed, the number of chemical
cleaning procedures for the main desalination stage RO1 nclp, y, unit is normally two
or three. The post-desalination stage RO2 does not have to be chemically cleaned as
often, usually only once or at most twice a year.

nclp,y ¼ nclp,y,unit  N units f:cleaning : ð5:252bÞ

nclp,y ¼ number of cleaning procedures in a year [no./year]


Nunits f. cleaning ¼ number of units of the RO tract to be cleaned []
nclp, y, unit ¼ number of cleaning procedures per year and unit [no./unit, y]
The permeate consumption for the membrane cleaning procedures is part of the
internal product water consumption of an SWRO plant and its percentage share is
calculated from VP,y and the plant’s total product water per year as shown in
Eq. (5.252d).

V P,y  100
V P,% ¼ : ð5:252dÞ
QPr,y

VP, % ¼ percentage of internal permeate consumption for cleaning of plant [%]


QPr, y ¼ annual product water amount [m3/year]
5.5 RO Process Units, Components and Overall System Configuration 703

The consumption of pretreated seawater for the cleaning procedures must be


taken into account when calculating the required performance of the RO system’s
pretreatment stages.

Wastewater Amount for Treatment


The volume of wastewater VWW,treatment,i generated during a cleaning sequence that
is contaminated with cleaning chemicals and has to be stored and treated in a waste
water system connected to the cleaning station results from the volume of the
cleaning solution Vcleaning, i drained in Step 8 and the volume VPF2,i generated during
permeate rinsing of the cleaning tank and membrane system in Step 9 of the cleaning
sequence (Eq. 5.253). The total wastewater volume VWW,treatment,clp generated by a
cleaning procedure is calculated from the volumes of the individual cleaning
sequences VWW,treatment,i (Eq. 5.253a). The total wastewater quantity VWW,treatm,τ
that is generated during the cleaning of an RO system is determined from the
wastewater volume of each cleaning procedure VWW,treatment,clp and the number of
procedures nclp,τ during the entire cleaning interval of an RO system (Eq. 5.253b).
The cleaning interval is taken to be the period τclean during which the number of
cleaning procedures nclp,τclean needed for cleaning all units Nunits f. cleaning of an RO
system just once are performed.

V WW,treatm,i ¼ V P9,i þ V cleaning,i : ð5:253Þ

X
i¼ncl,s

V WW,treatm,clp ¼ V WW,treatm,i : ð5:253aÞ


i¼1

i¼nclp,τclean
X
V WW,treatm,τclean ¼ V WW,treatm,clp : ð5:253bÞ
i¼1

VWW,treatment,i ¼ Wastewater volume to treatment during cleaning sequence [m3]


VWW,treatment,clp ¼ Wastewater volume to treatment during cleaning procedure
[m3]
nclp,τclean ¼ Number of cleaning procedures during cleanings interval time τclean
[]
VWW,treatm,τclean ¼ Wastewater volume for treatment during cleanings in cleanings
interval τclean [m3]

Chemicals Consumption
The chemical consumption required for membrane cleaning depends on the type and
extent of the fouling to be removed. However, the degree to which these membrane
coatings to be removed have aged also strongly impacts the intensity with which
cleaning is carried out and thus on the type of chemicals to be used as well as their
quantity and concentration. The conditions that give rise to fouling can change with
seawater quality and the efficiency of pretreatment and also with the mode of
704 5 Reverse Osmosis Membrane System: Core Process of SWRO

operation of the plant, i.e. the frequency of membrane cleaning affects the degree of
fouling and the consistency of the residues.
This means that in the design phase of a membrane desalination system, only a
rough estimate of chemical consumption for membrane cleaning is generally possi-
ble. For seawater desalination systems, it is evident that fouling as described above
consists of a mixture of predominantly organic/biological fractions together with
inorganic deposits of predominantly oxides and hydroxides of iron used as
flocculants in pretreatment. With fouling of this nature, if the organic/biological
content is not high, cleaning with citric acid alone or, for more extensive organic
fouling, a cleaning sequence comprising a combination of alkaline cleaning, possi-
bly with a complexing agent such as Na-EDTA and with the addition of an anionic
surfactant, followed by acid cleaning with a mineral acid or with citric acid, is
acceptable (see Table 5.30).
The chemical demand per cleaning sequence DCi ,X %,del ,cls1 is then calculated from
the quantities of the respective chemical components Ci in their supply concentration
X %,Ci ,del as they are used to prepare the necessary amount of cleaning solution V Ci ,cls1
and their specified solution concentration X %,Ci,solution (Eq. 5.254).

V Ci ,cls1 3 ρsolution  X %,Ci,solution


DCi ,X %,del ,cls1 = : ð5:254Þ
X %,Ci, del

DCi ,X %,del ,cls1 ¼ demand of chemical i at delivery concentration of X%,del used in


cleaning sequence i [kg]
V Ci ,cls1 ¼ cleaning solution volume for cleaning sequence i [l] [m3]
ρsolution ¼ density of cleaning solution [kg/l] [kg/m3]
X %,Ci,solution ¼ concentration of chemical i in cleaning solution [%]
X %,Ci ,del ¼ delivery concentration of chemical i [%]
The annual consumption C Ci X %,del ,y of a cleaning chemical Ci is then calculated by
multiplying its consumption per cleaning sequence DCi ,X %,del ,cls1 by the number of
cleaning sequences, i.e. cleaning procedures, per year ncls, y for which this chemical
is used (Eq. 5.254a).

C Ci X %,del ,y = DCi ,X %,del ,cls1 3 ncls,y : ð5:254aÞ

CCi X %,del ,y ¼ annual consumption of chemical i at delivery concentration [kg/year]


ncls, y ¼ annual number of cleaning sequences/procedures with chemical i [no./y]

Energy Consumption
The total energy consumed for cleaning an RO unit EC, cleaning, T is made up of the
energy consumption of the low-pressure flushing with the feed pump installation
E C,FF,SW,clpi at the beginning of the cleaning procedure and the service flush at
operation pressure of the RO system before the cleaned unit is put back into
operation, that of the cleaning pumps for permeate flushing E C,PF,clpi , and for
circulation of the cleaning solution EC,circ,clpi during the cleaning cycle as well as
5.5 RO Process Units, Components and Overall System Configuration 705

the energy consumption for heating the cleaning solution to its specified temperature
E C,heat,clpi . The total energy consumption for all cleaning procedures of an RO system
per year is calculated by multiplying the energy consumption of a single cleaning
procedure by their number carried out during each year nclp, y (Eq. 5.255).
 
E C,cleaning,T ¼ nclp,y  E C,FF,SW,clpi þ E C,PF,clpi þ E C,circ,clpi þ EC,heat,clpi : ð5:255Þ

EC, cleaning, T ¼ total cleaning energy consumption per year [kWh]


nclp, y ¼ number of cleaning procedures per year [no./y]
EC,FF,SW,clpi ¼ seawater flushing energy consumption per cleaning procedure
[kWh]
E C,PF,clpi ¼ permeate flushing energy consumption per cleaning procedure [kWh]
E C,circ,clpi ¼ recirculation energy consumption per cleaning procedure [kWh]
E C,heat,clpi ¼ cleaning solution heating energy consumption per cleaning
procedure
[kWh]
The energy consumption of the feed pump installation E C,FF,SW,clpi during a
cleaning procedure for low-pressure seawater flushing and for service flushing
done at the RO unit operating pressure is calculated from the respective flushing
flows F FF1,clpi and F FF2,clpi , the prevailing operating pressures pFP1,clpi and pFP2,clpi ,
the respective flushing durations τFF1,clpi and τFF2,clpi as well as the overall efficiency
ηPFT of the feed pumps with Eq. (5.255a).

F FF1,clpi ,  pFP1,clpi  τFF1,clpi þ F FF2,clpi ,  pFP2,clpi  τFF2,clpi


EC,FF,SW,clpi ¼ :
36  60  ηPFT
ð5:255aÞ

pFP1,clpi ; pFP2,clpi ¼ feed pressure during low-pressure seawater flushing [bar]


τFF1,clpi ; τFF2,clpi ¼ low-pressure seawater flushing duration [min]
The energy consumptions E C,PF,cls1 and E C,circ,cls1 during a cleaning cycle for the
permeate flushes to be carried out with the cleaning and flushing pumps of the
cleaning station and the circulation of the cleaning solution (Eqs. 5.255b and 5.255c)
are calculated in a similar way.

F PF1,cls1 ,  pFP1,cls1  τPF1,cls1 þ F PF2,cls1 ,  pFP2,cls1 τPF2,cls


EC,PF,cls1 ¼ 1
: ð5:255bÞ
36  60  ηPClT

F circ1,cls1 ,  pFP1,cls1  τcirc1,cls1 þ F circ2,cls1 ,  pFP2,cls1 τcirc2,cls


EC,circ,cls1 ¼ 1
:
36  60  ηPClFPT
ð5:255cÞ

E C,PF,cls1 ¼ permeate flushing energy consumption per cleaning cycle [kWh]


τPF1,cls1 ; τPF2,cls1 ¼ permeate flushing duration [min]
706 5 Reverse Osmosis Membrane System: Core Process of SWRO

pFP1,cls1 ; pFP2,cls1 ¼ cleaning/flushing pump operation pressure during permeate


flushing [bar]
E C,circ,cls1 ¼ recirculation energy consumption per cleaning sequence [kWh]
τcirc1,cls1 ; τcirc2,cls1 ¼ cleaning solution circulation duration [min]
pFP1,2,clpi ¼ cleaning/flushing pump operation pressure during cleaning sequence
[bar]
ηPClT ¼ efficiency of cleaning/flushing pumps []
If the cleaning procedure consists of more than one cleaning cycle, its energy
consumption is determined by the sum of the energy consumptions of permeate
rinsing and recycling of the cleaning solution in each cycle (Eqs. 5.255d and 5.255e).

Xcl,s
1¼n
E C,PF,clpi ¼ E C,PF,cls1 : ð5:255dÞ
i¼1

Xcl,s
1¼n
E C,circ,clpi ¼ E C,circ,cls1 : ð5:255eÞ
i¼1

ncl, s ¼ number of cleaning sequences per cleaning procedure []


The energy consumption for heating the cleaning solution to the temperature
specified for the cleaning process can be estimated from the volume of cleaning
solution V cleaning,cls1 , its density ρsolution,cls1 , and its specific heat capacity cp, solution,
together with the increase in temperature ΔT (Eq. 5.256). The empirical factor fC, heat
takes into account the additional energy consumption due to heat losses during the
cleaning process, especially during soaking. The value of this factor fC, heat can vary
between 1.2 and 1.5, depending on the dimensions of the cleaning tank, its
connecting piping, its location, and the outside temperatures in the cleaning station
and the RO plant. If soaking continues for a lengthy period, it may also be necessary
to completely reheat the cleaning system before restarting circulation of the cleaning
solution.

E C,heat,cls1 f C,heat  V cleaning,cls1  ρsolution,cls1  cp,solution  2, 778


 104  ΔT: ð5:256Þ

EC,heat,cls1 ¼ energy consumption for heating cleaning solution in cleaning


sequence [kWh]
V cleaning,cls1 ¼ cleaning volume in cleaning sequence [m3]
ρsolution,cls1 ¼ density of cleaning solution in cleaning sequence [kg/m3]
cp, solution ¼ specific heat capacity of cleaning solution [kJ/kg,K]
ΔT ¼ temperature increase [K]
fC, heat ¼ empirical factor for heat losses during the cleaning sequence/procedure
[]
If the cleaning procedure comprises several sequences, its total energy consump-
tion is the sum of their individual consumptions (Eq. 5.256a).
5.5 RO Process Units, Components and Overall System Configuration 707

Xcl,s
1¼n
EC,heat,clpi E C,heat,cls1 : ð5:256aÞ
i¼1

The power demand PD, heat for heating the cleaning solution in the cleaning
installation that is to be taken as the basis for dimensioning its heating equipment is
calculated using the same parameters as for the energy consumption, but additionally
considering the time specified for heating τheating and an empirical factor fD, heat that
takes into account the heat losses during the heating time as well as heating of the
mechanical components, like the tanks, pipes, etc., of the cleaning system
(Eq. 5.256b). This factor fD, heat can have a value of between 1.1 and 1.3 and depends
on the design of the cleaning system as well as the temperature conditions at the RO
system’s cleaning station.

PD,heat f D,heat
V cleaning,cls1  ρsolution,cls1  cp,solution  2, 778  104  ΔT
 :
τheating
ð5:256bÞ

PD, heat ¼ power demand for dimensioning of heating system [kW]


τheating ¼ heating time [h]
fD, heat ¼ empirical factor for heat losses during heating of the cleaning system []

5.5.2.5.5 Options for Improvement of Cleaning Efficiency


Gas Bubble Support During Cleaning
The effectiveness of a cleaning solution in removing deposits from a membrane
surface can be improved by pressurizing the solution with a gas released as bubbles
during the cleaning process. This not only produces increased and more uniform
turbulence on the membrane surface, but also improves the separation of solids from
the surface and their dispersion in the cleaning solution. The smaller the bubble size
during outgassing from the solution when microbubbles are formed, the better are
the results. The gaseous component of the cleaning solution may be air, which is
sucked in at the cleaning station using a venturi liquid jet pump and mixed with the
cleaning solution. In addition, the formation of microbubbles during the cleaning
process can be further promoted by employing special membrane cleaners
containing surface-active substances to influence the size range of the released
bubbles [147]. Instead of air, carbon dioxide can be added to the cleaning solution.
As a result, fouling deposits containing alkaline earth carbonates can also be
removed due to the chemical activity of CO2 [148]. But it has also been
demonstrated that the process with carbon dioxide is effective for dealing with
biological/organic fouling [149].
708 5 Reverse Osmosis Membrane System: Core Process of SWRO

Direct Osmosis: Osmotic Backflushing


Osmotic backflushing occurs when the osmotic pressure of the seawater and con-
centrate present on the feed side of the membrane system exceeds the system
operating pressure. The net driving pressure NDP then has a negative value and,
due to direct osmosis, permeate flows back through the membranes to the inlet side
of the membrane modules (see Sects. 5.1.5.1 and 5.1.5.1.1).
This operating situation arises when the pressure acting on a membrane unit is
relieved when it is taken out of service or before chemical cleaning commences. The
osmotic backflushing that then arises reduces the adhesion of fouling deposits on the
membrane surface and partially dissolves them so that they can be removed from the
system by subsequent flushing at low pressure (see Step 2 in Table 5.34). But for this
osmotic backflushing to be possible, sufficient permeate must be present to prevent
dehydration of the membranes (see Sects. 5.1.3.1 and 5.1.3.1.1). The resulting
additional volume on the feed side of the membrane system is discharged either on
the feed side or on the concentrate side of the membrane modules, whereby at the
same time low-pressure flushing is done to remove the existing fouling. For such
osmotic backflushing to be possible, the pipework must be appropriately configured
at the plant design stage.
Osmotic backflushing can also be done at the soaking phase of chemical cleaning.
For this purpose, the pressure on the membrane system is relieved and, thanks to the
correspondingly high ionic strength of the cleaning solution, this type of
backflushing on the permeate side can also be employed if the plant is so configured
to allow it.
The effect obtained by osmotic backflushing can also be achieved without
shutting down the membrane unit to be cleaned if a high-saline NaCl solution is
dosed into its feed. This increases the osmotic pressure in the water to be desalinated
to such an extent that it exceeds the operating pressure of the desalination plant. The
resulting negative net driving pressure, NDP, gives rise to a permeate backflow to the
feed side of the membranes during the dosing time of the highly concentrated
solution. With this so-called direct osmosis cleaning by high salinity solution
(DO-HS) process, the concentration of the highly saline dosing solution needed
for sufficient permeate backflow naturally depends on the osmotic pressure already
applied by the water to be desalinated and the resulting operating pressure of the RO
unit. With a relatively low salinity in the feed water of an RO unit, such as for
brackish water or water reuse effluent, a correspondingly negative NDP and perme-
ate backflow can be generated for the highly concentrated salt solution, even if the
dosing rate is quite low. Seawater desalination with its higher osmotic pressure of the
feed water and thus also of the associated operating pressure, however, requires a
significantly higher dosing capacity and concentration of the NaCl solution in order
to achieve a sufficiently high negative NDP and effective permeate backflushing
through direct osmosis. In seawater desalination plants, for example, a supersaline
concentration of the NaCl dosing solution of 17–25% is suggested with a dosing/
pulse time per cleaning step of approx. 20–60 s. Depending on the fouling potential
of the water to be desalinated, this direct osmotic cleaning process should be carried
out from once to several times per day [150, 151].
5.5 RO Process Units, Components and Overall System Configuration 709

Reverse Cleaning
For the chemical treatment process that is generally applied for RO membrane
systems, the cleaning solution flows through the membrane modules from the feed
side to the concentrate outlet, i.e. in the same direction as the water to be desalinated.
Since fouling products are preferably deposited at the frontmost elements of the
modules, this means that during cleaning the impurities removed there are
transported through all downstream elements to the concentrate outlet. For severe
fouling, it would therefore be advantageous to feed the cleaning solution to the
membrane modules from the concentrate side—reverse cleaning—as the fouling
deposits would then be discharged directly from the heavily contaminated front
elements without passing through the following, less heavily fouled elements. With
scaling deposits, their formation takes place primarily in the tail elements, so that the
conventional method of chemical cleaning is more advantageous for this type of
membrane coating [152].
So that these two options for admitting cleaning solution to the membrane
modules will be possible, provision must be made for this in the design of the
cleaning station and its connecting pipework.

5.5.2.5.6 Membrane System Disinfection and Membrane Preservation


Membrane System Disinfection
If biofouling occurs in RO membrane systems to such an extent that the above
described measures for physical and chemical cleaning of the membrane units
cannot remove a large remnant of the resulting deposits or their formation cannot
be adequately controlled, then additional inorganic and organic biocides can be used
to disinfect the membranes.
Treatment with biocides can be done:

• as an additional step in a chemical cleaning procedure during CIP treatment


• by continuous dosing during operation of the RO plant
• through intermittent dosing for a specified time at defined intervals during plant
operation.

Table 5.35 lists the inorganic and organic biocides commonly used in membrane
technology and how they are normally used for membrane disinfection. This table
also states indicative values for the concentration of the biocidal solution concerned
in the case of CIP treatment or for the dosing rate of the active substance in the case
of intermittent or continuous dosing, as well as the temperature and pH conditions
specified by the biocide manufacturers or the membrane manufacturers for their
application. The values for concentration and dosage rates are of course strongly
dependent on the biological activity of the RO feed water and the nutrients available
in the membrane system to support biological fouling, as well as on the degree of
fouling of the membranes when disinfecting and removing already existing biologi-
cally active deposits. Thus, for specific applications they may deviate considerably
from the values shown in the table.
Table 5.35 Biocides for disinfection and preservation of membranesa,b,c
Conditions of use
710

Abbreviation/ Concentration or dosing rate of Temperature


Biocide trade name Application active product pH ( C)
Inorganic biocides
Hydrogen peroxide/peracetic acid H2O2/PAA Disinfection 0.2% 3–4 <25
during CIP
Sodium bisulphite SBS Intermittent 500–1000 mg/l – 5–35
dosing
Preservation 0.5–1.5% >3
5

Organic biocides
Glutaraldehyde – Disinfection 0.1–1% 6.5– 5–35
during CIP 8.5
Preservation
Formaldehyde – Disinfection 0.1–1%
during CIP
Preservation
2,2-dibromo-3-nitrilopropionamide DBNPA Disinfection 30–50 ppm 4.5–
during CIP 8.5
Intermittent 10–30 mg/l
dosing
Continuous 5–20 mg/l
dosing
5-chlor-3-methyl-4-isothiazolin-3-on/3-methyl-4- Isothiazoline Intermittent 50–100 mg/l 2–9
isothiazolin-3-on CMIT/MIT dosing
Continuous 3–20 mg/l
dosing
Preservation 500–1000 ppm
a
DuPont Filmtec, “Tech Fact—Cleaning Procedures for FILMTEC FT30 Elements”
b
Hydranautics, “Technical Service Bulletin TSB 107—Foulants and Cleaning Procedures for Composite Polyamide RO Membrane Elements”
c
Reverse Osmosis Membrane System: Core Process of SWRO

Toray, “Operation, Maintenance and Handling Manual for Membrane Elements”—TMM-300 General instructions and conditions for RO cleaning—TMM-
310 Guidelines for RO cleaning—TMM-320 Instructions for chemical cleaning
5.5 RO Process Units, Components and Overall System Configuration 711

When used as a biocide for polyamide membranes, hydrogen peroxide mixed


with peracetic acid should only be used up to the total concentration of the
components as stated in the table and at a temperature below 25  C, otherwise the
membranes will be damaged. This mixture is normally only used for membrane
disinfection during CIP treatment and, prior to use, the membranes must be cleaned
by conventional alkaline and acidic treatment. Specifically, deposits should no
longer contain heavy metals as the oxidizing action of hydrogen peroxide is signifi-
cantly intensified catalytically in the presence of iron and other heavy metals.
With intermittent dosing, sodium bisulphite has a disinfecting action in the
specified concentration range. However, with this substance too, if heavy metals
and oxygen are also present, the oxidizing action on polyamide membranes will be
increased.
Of the organic biocides listed in Table 5.35, glutaraldehyde and formaldehyde are
only used in sequence for CIP treatment. The biocide DBNPA can be used for CIP
treatment as well as for intermittent or continuous dosing. If sodium bisulphite is
used at the same time, DBNPA is deactivated, which means that then 1.0–1.3 ppm
DBNPA must be added for each 1 ppm of SBS dosage. Empirical values for
intermittent dosing of DBNPA are dosing times of approx. 0.5–3 h every 2–5 days
depending on the severity of biological fouling.
DBNPA is more effective than the biocide isothiazoline. For membrane disinfec-
tion, this is used for both intermittent and continuous dosing. In the case of
intermittent dosing, a longer dosing and contact time is necessary due to its reduced
effectiveness, corresponding to dosing times of approx. 4–6 h every 2–5 days.
Although the retention rates for DBNPA and isothiazoline for brackish water and
seawater membranes are more than 98% and 99.5% respectively, neither biocide
should be dosed intermittently or continuously while the plant is in operation for
drinking water production. In this case, intermittent online disinfection is only
appropriate if the membrane system to be treated is flushed with seawater at low
pressure, during which time the permeate and concentrate are discarded. It is better to
use organic biocides for drinking water systems only for batchwise CIP treatment of
the membrane elements. In this case too, though, it is important to ensure that
flushing of the cleaning system and membranes with permeate will continue until
the organic biocides have been removed before the treated RO unit is returned to
service. This applies likewise to the biocides glutaraldehyde and formaldehyde when
these are used for CIP treatment.
The consumption of chemicals for continuous and intermittent dosing of biocides
can be calculated on the basis of the equations listed under Sect. 5.5.2.4,
Eqs. (5.241–5.241c, 5.242–5.242a, 5.245a–5.245e) while the power demand of the
dosing equipment may be determined on the basis of Eqs. (5.246–5.246b). The
water, chemical, and energy consumptions resulting from additional disinfection
sequences during CIP treatment of a membrane unit are determined as set out in
Sect. 5.5.2.5.4.
712 5 Reverse Osmosis Membrane System: Core Process of SWRO

Membrane Preservation During Plant Shutdown


If an RO system is taken out of operation, the first step after shutting it down is to
expel the concentrate from the system by low-pressure flushing with feed water.
What further measures are taken depends on how long the system is to be shut down,
i.e. whether it will be returned to service after a short outage time or if it is to be shut
down for a lengthy period.

Short-Term Shutdown Depending on manufacturer, a period in which the mem-


brane unit is out of operation for between 2 and 7 days is regarded as a short-term
shutdown. During this period, it is considered sufficient to carry out low-pressure
flushing of the membrane system every 24–48 h to prevent a loss of performance of
the membranes due to biological growth in the elements. One of the membrane
manufacturers specifies a low-pressure backflush only every 5 days and a permissi-
ble time for the short shutdown not to exceed 30 days.
In the case of a seawater desalination system, it is in order to carry out the
low-pressure flushing with low-saline permeate or product water as this also serves
to protect the metallic pipes and the RO feed pump installation against corrosion
even during shutdowns lasting several days or longer. In this case, for polyamide
membranes a non-oxidizing biocide can be added to the rinsing water, whereas for
CTA membranes chlorine at a concentration of up to 0.3 mg/l Cl2 can be added.

Long-Term Shutdown If the time during which the membrane unit is out of
operation lasts longer than the timeframe specified above for the short-term shut-
down, the membranes have to be preserved to prevent biological growth.
Inorganic biocides, like sodium bisulphite (SBS) as well as organic biocides,
e.g. glutaraldehyde, formaldehyde, or isothiazoline, can be used for preservation.
The respective concentration of the biocides and their conditions of use during
preservation are listed in Table 5.35. Of all the biocides shown here, however,
predominantly sodium bisulphite, SBS, at a concentration of 0.5–1.5% is used in
practice. If glutaraldehyde or formaldehyde is used, the permeate flux of the mem-
brane elements can be reduced. Although DBNPA is a very effective short-term
biocide, it is not effective over a lengthy period as at a pH of 7 it decomposes into
carbon dioxide, ammonium, and bromide within about 24 h, so it cannot be used for
preservation during long-term shutdowns. Isothiazoline is sometimes used for con-
serving CTA membrane elements.
Prior to the conservation measure, the membrane unit is chemically cleaned as
described in Sect. 5.5.2.5.3, Steps 1–9, as shown in Table 5.34. The system is then
filled with the preservation solution in accordance with Steps 1, 4, and 5 of
Table 5.34. During circulation of this solution (Step 5), the concentration can be
adjusted as needed by adding further preservation agent.
If SBS is used, the pH of the preservation solution must be monitored. Oxidation
of the SBS with oxygen lowers the pH of the solution, and before it reaches a value of
3, for polyamide membranes the preservation solution must be renewed. If pH
monitoring during preservation is not possible, at a storage temperature of up to
5.5 RO Process Units, Components and Overall System Configuration 713

27  C the preservation solution for this type of membrane should be replaced every
30 days and every 15 days at higher temperatures.
When preserving CTA membranes with sodium bisulphite, sodium benzoate,
SBA, is added to the preservation solution in addition to the SBS, and its pH adjusted
to 4.5. The pH should not fall below 4.0 during the shutdown period. The preserva-
tion solution should be changed before, but at least every 3 months.
When the plant is to be returned to operation, the membrane system must be
flushed with permeate beforehand until the preservation solution is completely
ejected from the system (Step 9 in Table 5.34). Then operation can be resumed in
accordance with Steps 10 and 11.
The additional consumptions of water, chemicals, and energy during short- and
long-term membrane unit shutdowns and also the amount of waste water to be
treated are calculated as described for chemical cleaning of the membranes in
Sect. 5.5.2.5.4.

5.5.2.6 RO Membrane Unit and System Configuration

5.5.2.6.1 Membrane Unit/Array Configuration


Array Capacity and Number of Arrays
RO systems have a modular configuration. For membrane desalination plants in the
lower capacity range, the desired product performance is achieved by membrane
modules connected in parallel, whereas for larger RO plants the target capacity is
achieved by membrane units or arrays consisting of multiple membrane modules,
these being likewise connected in parallel.
The maximum product flow that is possible with such a combination of mem-
brane modules or arrays or trains is determined by a number of factors. This is first
the available delivery of the feed pump installation and, in the case of seawater
desalination systems, the type and concentrate flow capacity of the selected energy
recovery system.
For an ERD directly connected on the HP pump side, i.e. a Pelton turbine or a
turbocharger, for which the delivery of the HP pump system FP, RO also corresponds
to the feed flow FF, RO of the RO array (FP, RO ¼ FF, RO), the maximum capacity of
an array CPr, array, max is calculated from the delivery capacity of the feed pump
installation FP, RO, avail using Eq. (5.257).

CPr,array, max ¼ F P,RO,avail  24  Y RO1 : ð5:257Þ

CPr, array, max ¼ max. product capacity of array at available feed pump flow [m3/
day]
FP, RO, avail ¼ available feed pump flow [m3/h]
For an isobaric energy recovery system in which the high-pressure partial flow
from the ERD is independently fed to the RO array and in which the delivery of the
HP feed pump installation is approximately equal to the RO product flow (FP,
RO ffi FF, RO  YRO), Eq. (5.257a) applies to determine the maximum product
capacity CPr, array, max of an array.
714 5 Reverse Osmosis Membrane System: Core Process of SWRO

C Pr,array, max ffi F P,RO,avail  24: ð5:257aÞ

With a directly coupled ERD, i.e. Pelton turbine and turbocharger, the available
concentrate flow of the energy recovery system may determine the maximum flow
capacity of the membrane array and not the delivery capacity of the HP pump. In this
case, the array’s maximum flow capacity is calculated according to Eq. (5.257b).

Y RO1
C Pr,array,max ¼ F C,EDR,avail  24  : ð5:257bÞ
1  Y RO1
FC, EDR, avail ¼ available concentrate flow rate of ERD [m3/h]
YRO1 ¼ product recovery rate of RO1 []
A further criterion for dimensioning the product capacity of an array is the
bandwidth in which the entire product flow of the RO system is to be provided.
The product flow of an array on its own can only be varied to a limited extent within
a range of around 10–15%. By appropriate dimensioning of the product capacities of
the individual arrays of the RO system and thus also of the number of arrays, and by
alternately putting and taking individual arrays into and out of operation, the
flexibility of the RO flow rate as a whole can be correspondingly enhanced. Thus,
the product capacity of an individual array can be dimensioned in line with the
specified minimum rC, min of the total RO capacity that is to be made available
(Eq. 5.257c).

C Pr,RO,nom:  r C, min
C Pr,array, min ¼ : ð5:257cÞ
100
C Pr,RO, min :  100
r C, min ¼ :
C Pr,RO,nom:

CPr, array, min ¼ product capacity of array for minimum RO capacity [m3/day]
CPr, RO, nom. ¼ nominal output capacity of RO system [m3/day]
CPr, RO, min . ¼ minimum output of RO system [m3/day]
rC, min ¼ percentage minimum output to be provided by the RO system [%]
The RO system’s total output is then varied in accordance with this array capacity
or a multiple of it. If the output flow is to be changed in even smaller increments,
i.e. if the flexibility of the RO system is to be even greater, the capacity of the
individual arrays must be further reduced and their number increased accordingly.
The number of arrays Narray results from the RO system’s nominal product capacity
CPr, RO, nom. and the product capacity CPr, array selected for its membrane units or
arrays (Eq. 5.258).

C Pr,RO,nom:
N array ¼ : ð5:258Þ
C Pr,array

Narray ¼ number of arrays or trains of the RO system []


5.5 RO Process Units, Components and Overall System Configuration 715

For large SWRO plants, the array/train capacity is in a range from 10,000 m3/day
up to 25,000–30,000 m3/day.

Standby Arrays and Overloading


Thanks to the modular design of an RO system, routine maintenance such as
cleaning and replacement of the membranes can be done successively for the
individual arrays. This also applies to servicing of the system’s mechanical
components. If the RO system is required to maintain its product flow without
reduction or interruption, any array shutdowns and associated capacity losses must
be compensated or prevented from occurring.
To avoid such output losses, from the viewpoint of an RO system there are two
possibilities:

• Within the RO array configuration, an additional standby unit is provided, which


is operated during maintenance of one of the other arrays. In this case, however, if
such a unit is out of service for a prolonged period, measures to prevent biological
fouling of the membranes must be taken as described above in Sect. 5.5.2.5.6.
• During servicing or maintenance, the output of the units that are taken out of
operation is made good by the arrays that remain in operation by overloading
them. However, the average product flux JP, S ∅ , A, nom of the remaining
membrane units (see Sect. 5.2.2, Table 5.12) should either not be exceeded at
all or only exceeded to a certain degree as allowed by the membrane manufacturer
for the duration of the maintenance work. The extent to which the value of the
average product flux increases during overloading depends on the total number of
arrays Narry, nom of the RO system and the number of units Narry, maint remaining in
operation during maintenance (Eq. 5.259).

N arry,nom
J P,S∅,A,overl ¼ J P,S∅,A,nom  : ð5:259Þ
N arry,maint

JP,S∅,A,overl ¼ average specific product flux of arrays during overloading [l/m2h]


JP,S∅,A,nom ¼ nominal average specific product flux of arrays [l/m2h]
Narry, nom ¼ number of arrays during nominal RO operation []
Narry, maint ¼ number of arrays during maintenance operation of RO []

To cope with this overloading, the RO feed pump system of the RO is to be


dimensioned in such a way that the increased operating pressure required for the
higher value of JP,S∅,A,overl is made available. In addition, the arrays must be
designed to allow the higher flow rates of feed, product, and concentrate.
For the overloading it is necessary that the feed pump system of the reverse
osmosis is dimensioned in such a way that the higher operating pressure required for
the higher value of JP,S,∅,A,overl is made available. In addition, the arrays must be
designed to allow the higher flow rates of feed, product, and concentrate.
716 5 Reverse Osmosis Membrane System: Core Process of SWRO

As Eq. (5.259) shows, overloading is best performed when the number of arrays
is relatively high, i.e. the array’s flow dimensioning is not too close to the values set
by the available pump capacity or the ERD’s concentrate flow. It is also apparent that
this manner of compensation for performance losses during maintenance is only
possible with larger desalination plants, which also have a correspondingly higher
number of arrays. In contrast, for smaller RO systems with fewer arrays, the amount
by which JP, S ∅ , A, nom is exceeded is too high and a standby unit is required to
compensate for the output losses.

5.5.2.6.2 RO System Configuration


The configuration of the main desalination unit RO1 of a seawater RO system is
essentially determined by the type of energy recovery device selected for it.
There are three options for RO1 configuration, namely:

1. The train configuration—the main desalination stage is made up of a number of


trains, each of which is equipped with its own feed pump installation, comprising
supply pump, cartridge filter, HP pump and, if applicable, ERD supply pump; a
membrane array; and a dedicated energy recovery system.
2. The feed centre configuration—several membrane arrays are assigned to a
lesser number of feed pump installation groups and are fed from these via a
common header system. Each membrane array, though, has its own energy
recovery system.
3. The three-centre configuration—on the feed side, this is structured like the feed
centre configuration. However, the energy recovery devices are combined into a
single energy recovery centre of modular design which is supplied with concen-
trate from the arrays of the membrane centre via a concentrate ring line. The ERD
high-pressure supply generated in the ERD centre is fed into the common feed
header of the feed centre.

Graphics of the three RO system configurations in simplified form are shown in


Fig. 5.100.
The pump symbols in this graph stand for the complete feed pump installations,
that is in the train configuration the supply pump, cartridge filter, HP pump, and
possibly the ERD supply pump of the train. For the feed centre and the three-centre
configurations, these are the respective assemblies of supply pumps, cartridge filters,
ERD supply pumps, and HP pumps.
The train configuration is an option for all energy recovery systems described
under Sect. 5.5.2.2.3. Energy recovery systems directly coupled to the HP pump,
though, such as the Pelton turbine, are limited solely to the train configuration.
However, with isobaric energy recovery, where the high-pressure supply from the
ERD is decoupled from the feed on the HP pump side, this also allows the use of the
other two configuration options.
5.5 RO Process Units, Components and Overall System Configuration 717

Train configuration
Feed supply
systems
RO

ERD

RO

Feed ERD
Permeate

RO

ERD

RO
Concentrate
ERD

Feed center configuration

RO

ERD
Feed center

RO

Feed
ERD
Permeate

RO

ERD

RO
Concentrate
ERD

Three center configuration

RO

Feed center
RO

Permeate
Feed
RO

RO

ERD
Concentrate

Energy recovery center

Fig. 5.100 RO configuration options


718 5 Reverse Osmosis Membrane System: Core Process of SWRO

In the train configuration, the delivery of the pumps in its feed pump installation
and also the performance of the energy recovery device connected to the train is
determined by the product capacity selected for the train array (Eqs. 5.257, 5.257a
and 5.257b). But with the feed centre and three-centre configurations, it is possible to
optimize and specify the delivery of the feed pumps independently of the product
flow capacity of the arrays as a fixed number of arrays is no longer assigned the same
number of feed pumps as is the case with the train configuration.
The total delivery capacity F FHP ,centre of the HP pumps to be made available in the feed
centre of a feed centre configuration with isobaric energy recovery is almost equivalent to
the product flow capacity FPr, RO1 of the RO system. As Eq. (5.260) shows, the product
capacity CPr, array of an array at a given RO product flow FPr, RO1 depends on the number
of arrays Narray, centre in the membrane centre of the configuration.

C Pr,array  N array,centre
F FHP ,centre ffi F Pr,RO1 ffi : ð5:260Þ
24
F FHP ,centre ¼ high-pressure feed flow to feed centre [m3/h]
CPr, array ¼ product capacity of array [m3/day]
Narray, centre ¼ number of arrays of centre []
The delivery rating F PHP ,centre of each HP pump in the feed centre is calculated
from the total product flow CPr, RO1 of the RO1 stage and the number N PHP ,centre of
high-pressure pumps installed there (Eq. 5.260a).

CPr,array  N array,centre CPr,RO1


F PHP ,centre ffi ffi : ð5:260aÞ
24  re 24  N PHP ,centre

F PHP ,centre ¼ delivery of feed centre HP pumps [m3/h]


N PHP ,centre ¼ number of HP feed pumps in feed centre []
CPr, RO1 ¼ product capacity of RO1 [m3/day]
The total flow F F ERD ,centre to be provided by the feed centre to the energy recovery
centre is calculated using Eq. (5.260b). F FERD ,centre also largely corresponds to the
high-pressure partial flow F D,HPERD ,centre supplied from the ERD centre to the feed
line to the membrane centre and the concentrate flow FC, ERD centre of the isobaric
energy recovery system to be installed in the ERD centre. For details of the calcula-
tion of isobaric energy recovery systems, see Sect. 5.5.2.2.3.

CPr,RO1 1
F FERD ,centre ffi  1 ffi F D,HPERD ,centre ffi F C,ERD centre : ð5:260bÞ
24 Y RO1

F F ERD ,centre ¼ feed to ERD centre [m3/h]


F D,HPERD ,centre ¼ high-pressure discharge flow of ERD centre [m3/h]
YRO1 ¼ recovery factor of RO1 []
FC, ERD centre ¼ concentrate flowrate of ERD centre [m3/h]
The delivery F PsupplyERD ,centre of each ERD supply pump in the feed centre is
calculated from the total product flow CPr, RO1 and the number of pumps
N Psupply ERD ,centre in the centre with Eq. (5.260c).
5.5 RO Process Units, Components and Overall System Configuration 719


C Pr,RO1 1
F PsupplyERD ,centre ffi  1 : ð5:260cÞ
24  N Psupply ERD ,centre Y RO1

F PsupplyERD ,centre ¼ ERD supply pump discharge of feed centre [m3/h]


N Psupply ERD ,centre ¼ number of ERD supply pumps of feed centre []
The relationships shown above enable both pump-side optimization with regard
to pump efficiency aspects in the feed centre and the determination of an economic
and operational optimum for the size of the arrays of the membrane centre of a centre
configuration [153].
The efficiency of centrifugal pumps depends on their specific speed and delivery.
For a defined specific speed, the efficiency increases with increasing delivery rate.
The extent to which the efficiency of a centrifugal pump is influenced by its delivery
rate, though, also depends on its design. For more details, see Sect. 5.5.2.2.2,
Figs. 5.74 and 5.75, Eqs. 5.221–5.221b. By optimizing the centrifugal pumps of
the feed centre of a centre configuration taking these aspects into account, the
efficiencies of the HP and supply pumps can be appreciably improved. This optimi-
zation is achieved by correspondingly adjusting the number of pumps in the feed
centre and thus changing the delivery of the respective pump type and checking the
resulting efficiency of the pumps.
Optimization of the array size examines the question of flexibility in bridging
shutdowns due to membrane replacement and membrane cleaning either by array
overloading or operating standby arrays and in changing the plant’s product flow
capacity, also the size of the membrane cleaning system as well as the cost of piping
and fittings for the arrays. The centre design therefore makes it possible to combine
an efficiency-optimized pump design in the feed centre with arrays whose size
corresponds to an optimum in plant flexibility for bridging capacity losses during
shutdown as well as under consideration of cost aspects.
In order to achieve a uniform distribution of the HP feeds from the feed centre and
the ERD centre via the common HP header of a centre configuration to the individual
arrays in the membrane centre, the feed line of each array must be equipped with
flow control equipment consisting of a control valve with associated feed and
pressure control loop. This control loop, which allows the feed rate to be adapted
to match the specific operating conditions of each array, is necessary because in
practice membrane permeabilities and pressure losses in the individual arrays will
not be the same but may differ considerably. For this reason, the feed rate and
operating pressure have to be individually adjusted for each array. The operating
pressure to be maintained in the feed centre by the HP and ERD supply pumps is
then oriented to those arrays of the membrane centre with the most unfavourable
operating conditions, i.e. those with the lowest performance. For the other arrays that
require a lower operating pressure, the pressure must then be throttled down in the
feed lines to these arrays. It is thus possible that the advantages obtained by
optimizing the efficiency of the pumps of the feed centre will be partially negated
by this pressure throttling in the array feed lines of the membrane centre.
720 5 Reverse Osmosis Membrane System: Core Process of SWRO

While in a centre configuration a combination of speed control of the pumps of


the feed centre and feed control of the individual arrays is necessary for supply of the
membrane centre, in a train configuration the feed and pressure can be controlled
more easily and individually for each train solely via the speed control of the feed
pumps of the train.
The feed centre configuration is to be considered in particular for SWRO plants
with medium to high product flow capacities and a number of RO desalination plants
have been realized with this configuration.
The three-centre configuration is particularly suitable for very large RO seawater
desalination plants whose product flow has to cover a wide range with a high degree
of flexibility. So far, some large-scale SWRO plants have been built in Israel with
this configuration at Ashkelon, Hadera, and Sorek.
The most common configuration of membrane units for reverse osmosis seawater
desalination is the train configuration. Throughout the performance range from low
to high desalination capacity, it does not matter whether centrifugal or positive
displacement HP pumps are installed in the feed pump installation and any type of
energy recovery system may be used. Further advantages of this system configura-
tion are first its high flexibility and second that the feed to each train can be
individually adjusted to match its operating conditions.
Annexes

Annex 5.A1
Annexes

Table 5.36 Reverse osmosis seawater and brackish water membrane manufacturer

Head Software
Manufacturer Company quarter Homepage links Design Normalization Software links
Polyamide—TFC—Spiral wound membranes
CSM Toray South http://www.csmfilter.com/ CSMPRO CSMNORM http://www.csmfilter.com/
Chemical Korea
Korea
DuPont DuPont Water USA https://www.dupont.com/water/ WAVE FTNORM https://www.dupont.com/water/
Solutions— technologies/reverse-osmosis-ro. System data resources/design-software.html
Filmtec html normalization https://water.custhelp.com/app/
answers/detail/a_id/478/kw/
normalization
Hydranautics Hydranautics USA/ https://membranes.com/solutions/ IMSDesign ROdata XL https://membranes.com/solutions/
Nitto Group Japan products/ro/ software/
Company
Koch Koch USA https://www.kochmembrane.com/ RoPro – –
Separation en-US/products/spiral-
solutions membranes/fluid-systems-
elements
Lewabrane LANXESS Germany https://lpt.lanxess.com/en/ LEWAPLUS https://lpt.lanxess.com/lewaplus-
products-lpt/product-groups/ software/
reverse-osmosis/
Membranium Membranium Russia https://www.membranium.com/ NanoTech – https://www.membranium.com/
catalog/spiral-wound-membrane- PRO technical-info/software/
elements/reverse-osmosis/
Suez Suez Water USA/ https://www. Winflows – https://www.
Technologies France suezwatertechnologies.com/ suezwatertechnologies.com/
721

& Solutions products/spiral-wound- resources/winflows


membranes
(continued)
Table 5.36 (continued)
722

Head Software
Manufacturer Company quarter Homepage links Design Normalization Software links
Toray Toray Japan http://www.toraywater.com/ Toray DS2 Toray Trak https://ap3.toray.co.jp/toraywater/
Membrane Industries Inc. index.html
Polyamide—TFC—Nano—Spiral wound membranes
LG chem LG Water South http://www.lgwatersolutions. Q+ Q See http://www.lgwatersolutions.com/
Solutions Korea com/ Projection Normalization en/tools/software
NanoH2O Software software
5

Cellulose triacetate—Hollow fiber membranes


Toyobo Toyobo Co, Japan http://www.toyobo-global.com/ HOLLOSEP – –
Ltd seihin/ro/index.htm
Reverse Osmosis Membrane System: Core Process of SWRO
Annex 5.A2

Table 5.37 Characteristics of RO membrane elements for seawater desalination—polyamide spiral wound composite
Seawater RO membrane elements
Annexes

Type of RO membranes Polyamide spiral wound composite


Membrane element category High
Membrane element parameters Symbol Unit Normal rejection High flux/low energy Low fouling
Element dimensions
• Diameter/length inch 800 1600 800 800 1600 800
7.9/40 15.8/40 7.9/40 7.9/40 15.8/40 7.9/40
mm 201/1016 402/1016 201/1016 201/1016 402/1016 201/1016
• Feed spacer mil 28–34 28 28–34 28–34 28 34
mm 0.71–0.86 0.71 0.71–0.86 0.71–0.86 0.71 0.86
• Membrane area SE ft2 380–440 1700 - 1760 400–440 370–440 1700-1760 370–440
m2 35.3–40.9 158–163.5 37.2–40.9 34.4–40.9 158–163.5 34.4–40.9
Standard test conditions
• Pressure pFE(St) bar 55.2

• Temperature tFE(St) C 25
• NaCl—concentration cFE,NaCl mg/l 32,000
(St)
• Boron concentration/pH cFE,B(St) mg/l/pH 5/8
• Recovery rate of element YE(St) % 8/10a 8/10a 8/10a 8/10a 8 8/10a
Element performance at standard test conditions
• Capacity max. CPE(St) m3/day 23–31.2 92.4–136 21.9–27.3 28.0–50.0 110–150 24.6–45.5
• Permeate flow max. FPE (St) m3/h 0.96–1.3 3.85–5.67 0.91–1.14 1.17–2.08 4.58–6.25 1.03–1.90
• Permeate flux max. JPE(St) m3/m2, h 0.0272– 0.0244– 0.0245– 0.0339– 0.0290– 0.0298–
0.03178 0.0347 0.0277 0.0509 0.0382 0.0464
l/m2, h 27.20–31.78 24.36–34.68 24.53–27.74 33.92–50.94 29.0–38.23 29.80–46.35
(continued)
723
Table 5.37 (continued)
724

Seawater RO membrane elements


Type of RO membranes Polyamide spiral wound composite
Membrane element category High
Membrane element parameters Symbol Unit Normal rejection High flux/low energy Low fouling
• Salt (NaCl) rejection stabilized RENaCl(St) % 99.70–99.80 99.75–99.80 99.80–99.86 99.70–99.80 99.60–99.80 99.80
nominal.
• Boron rejection nominal. at pH REB(St) %/pH 92–95/8 92–95/8 93–96/8 89–95/8 92–95/8 92/8
91–92/6.5–7 91–93/6.5–7
a
5

Recovery rate of hydranautics


Reverse Osmosis Membrane System: Core Process of SWRO
Annexes 725

Annex 5.A2.1

Table 5.38 Characteristics of RO membrane elements for seawater desalination—polyamide


spiral-wound nanocomposite and cellulose triacetate (CTA)—hollow fiber
Seawater RO membrane elements
Polyamide spiral wound
Type of RO membranes nanocomposite CTA—
Membrane element category High Hollow
Membrane element parameters Symbol Unit rejection High flux fiber
Element dimensions
• Diameter/length inch 800 800 1100
7.9/40 7.9/40 11/81.8
mm 201/1016 201/1016 280/2078
• Feed spacer mil 28–34 28–34 –
mm 0.71–0.86 0.71–0.86
• Membrane area SE ft2 400–440 400–440 –
m2 37.2–40.9 37.2–40.9 ~1200
Standard test conditions
• Pressure pFE(St) bar 55.2 53.9

• Temperature tFE(St) C 25
• NaCl—concentration cFE,NaCl(St) mg/l 32,000 35,000
• Boron concentration/pH cFE,B(St) mg/l/pH 5/8 –
• Recovery rate of element YE(St) % 8 8 30
Element performance at standard test conditions
• Capacity max. CPE(St) m3/day 22.7–37.5 51.9–57.0 47.5
• Permeate flow max. FPE (St) m3/h 0.95–1.56 2.16–2.38 1.98
• Permeate flux max. JPE(St) m3/m2, h 0.0255– 0.0581– 0.00165
0.0381 0.0582
l/m2, h 25.5–38.1 58.1–58.2 1.65
• Salt (NaCl) rejection RENaCl(St) % 99.85– 99.80 99.6
stabilized nominal 99.89
• Boron rejection nominal at pH REB(St) %/pH 93/8 89/8 –
Annex 5.A3

Table 5.39 Characteristics of brackish water RO membrane elements used in 2. Pass of RO seawater desalination—Polyamide spiral wound and polyamide spiral-
726

wound nanocomposite
Brackish water RO membrane elements
Polyamide spiral wound
Type of RO membranes Polyamide spiral wound composite nanocomposite
Membrane element category
Membrane element High High High flux/lower
parameters Symbol Unit Normal rejection High flux/lower salt rejection rejection salt rejection
Element dimensions
5

• Diameter/length inch 800 1600 800 800 1600 800 800


7.9/40 15.8/40 7.9/40 7.9/40 15.8/40 7.9/40 7.9/40
mm 201/ 402/1016 201/ 201/1016 402/ 201/ 201/1016
1016 1016 1016 1016
• Feed spacer mil 28–34 28 28–34 28–31 28 28–34 28–34
mm 0.71– 0.71 0.71– 0.71–0.79 0.71 0.71– 0.71–0.86
0.86 0.86 0.86
• Membrane area SE ft2 380–440 1600– 380–440 400–440 1700 400–440 400–440
1700–
1760
m2 34–40.9 148.6– 35.3– 37.2–40.9 158 37.2– 37.2–40.9
158–163.5 40.9 40.9
Standard test conditions
• Pressure pFE(St) bar 15.5 10.3 15.5 8.6–10.3

• Temperature tFE(St) C 25
• NaCl— cFE,NaCl(St) mg/l 1500a/2000
concentration
• Boron cFE,B(St) mg/l/pH 5/8–10 5/10
concentration/pH
Reverse Osmosis Membrane System: Core Process of SWRO
• Recovery rate of YE(St) % 15
element
Annexes

Element performance at standard test conditions


• Capacity max. CPE(St) m3/day 36–48 155–169 22.7– 32.2–49.2 128.7– 39.7– 39.7–47.9
45.8 155.2 47.9
• Permeate flow FPE (St) m3/h 1.50–2.0 6.46–7.04 0.95– 1.34–2.05 5.36– 1.65– 1.65–1.996
max. 1.91 6.47 1.996
• Permeate flux JPE(St) m3/m2, h 0.0441– 0.0435– 0.0268– 0.0361–0.0501 0.0339– 0.0445– 0.0445–0.0488
max. 0.0489 0.0429 0.0467 0.0409 0.0488
l/m2, h 44.12– 42.94– 26.79– 36.07–50.12 33.94– 44.47– 44.47–48.80
48.90 43.52 46.66 40.93 48.80
• Salt (NaCl) RENaCl(St) % 99.5– 99.6–99.7 99.75– 99.2–99.7 99.3– 99.60– 99.0–99.6
rejection nominal 99.7 99.80 99.6 99.78
• Boron rejection REB(St) %/pH – – – 68/8–83/8–93/10– 96/10 – –
nominal at pH 95/10–96/10
a
Hydranautics and CSM
727
728 5 Reverse Osmosis Membrane System: Core Process of SWRO

5.A4 Reference Seawater Feed-PHREEQC-Pitzer Calculation

Please find this electronic additional material online under sn.pub/extras (See Docu-
ment 5.1)

5.A5 Reference Seawater-First Pass Concentrate-PHREEQC-Pitzer


Calculation-without Acid Dosing

Please find this electronic additional material online under sn.pub/extras (See Docu-
ment 5.2)

5.A6 Reference SeawaterSecond Pass Concentrate-PHREEQC:


WATEC4 Calculation

Please find this electronic additional material online under sn.pub/extras (See Docu-
ment 5.3)

5.A7 Overview Report of a Membrane Design Calculation of a 2-Pass


Seawater Desalination (Data Source: Toray DS2 Membrane Design
Software)

Project SWRO reference plant; Seawater TDS ¼ 35,000 mg/l;


100,000 m3/day; Feed temperature 25  C;
Case 2 stage; Y1 ¼ 45%, Y2 ¼ 90%, H2SO4 dosing 1. Pass;
NaOH dosing 2. pass; AMLT ¼ 3 years
Feed water type Sea open, conventional pretreatment

Parameters Units Overall Pass 1 Pass 2


Raw water TDS mg/l 36,058.5 34,558.9 280.76
Feed EC @25C uS 52,658 50,914 582
Feed pressure bar 0.0 50.72 11.74

Temperature C 25
Total DP bar 5.313 0.603 4.709
Brine pressure bar 50.12 50.11 7.029
Fouling (max) 3.00 years 0.850 0.955
SP % increase (max) 3.00 years 22.50% 22.50%
Recovery % 42.4% 45.0% 90.0%
Feed flow m3/day 235,801 246,911 111,110
Recycle flow m3/day 11,110 11,110
Product flow m3/day 100,000 111,110 100,000
(continued)
Annexes 729

Parameters Units Overall Pass 1 Pass 2


Average flux l/m2/h n/a 13.005 35.02
Concentrate flow m3/day 135,801 135,801 11,110
Product TDS mg/l 7.51 270.0 7.51
Concentrate TDS mg/l 62,612 62,612 2737

Flow diagram

Feed to Net Feed to Concentrate Product


system Pass 1 Pass 1 Pass 1
Ions Units 4 10 30 14
Ca mg/l 423.2 404.5 734.7 0.944
Mg mg/l 1316 1258 2285 2.935
Na mg/l 11,067 10,612 19,217 95.35
K mg/l 409.1 392.8 710.2 4.934
Ba mg/l 0.110 0.105 0.191 0.0002
Sr mg/l 8.150 7.791 14.150 0.018
HCO3 mg/l 107.5 102.2 187.0 1.270
CO3 mg/l 3.705 1.337 1.463 4.45E-05
CO2 mg/l 0.914 5.340 5.013 4.948
Cl mg/l 19,866 19,039 34,489 155.8
SO4 mg/l 2781 2666 4841 6.942
NO3 mg/l 2.270 2.181 3.940 0.0307
F mg/l 1.330 1.278 2.309 0.018
Br mg/l 68.97 66.10 119.7 0.541
(continued)
730 5 Reverse Osmosis Membrane System: Core Process of SWRO

Feed to Net Feed to Concentrate Product


system Pass 1 Pass 1 Pass 1
Ions Units 4 10 30 14
SiO2 mg/l 0.700 0.671 1.215 0.0068
B(boron) mg/l 4.000 4.146 6.538 1.223
TDS mg/l 36,059 34,559 62,612 270.0
Conductivity @25C μS 52.659 50.914 90,406 558.4
pH pH 8.10 7.3 7.58 5.5
Osmotic pressure bar 26.593/ 25.51/24.84 45.665/ 0.222/
(DS1/Pitzer) 25.94 46.37 0.22
LSI/SDSI 0.92/0.09 0.08/0.74 1.09/0.12
CaSO4/SrSO4% % 22.5%/ 21.3%/ 45.6%/
14.4% 13.6% 34.3%
BaSO4/SiO2% % 428.4%/ 407.3%/ 686.1%/1%
0.5% 0.5%
Pitzer % solubility Calcite/ 187%/ 33%/271% 174%/
dolomite 8626% 7828%
Pitzer % solubility CaSO4/ 23%/23% 22%/22% 43%/43%
SrSO4

Concentrate Pass Product Pass


Feed to Pass 2 2 2
Ions Units 40 26 20
Ca mg/l 0.944 9.367 0.0069
Mg mg/l 2.935 29.13 0.0214
Na mg/l 95.35 969.4 2.512
K mg/l 4.934 47.60 0.188
Ba mg/l 0.0002 0.0024 1.79E-06
Sr mg/l 0.018 0.180 0.0001
HCO3 mg/l 6.783 61.35 0.206
CO3 mg/l 1.320 17.623 0.006
CO2 mg/l 0.0033 0.0242 0.0066
Cl mg/l 155.8 1521 3.930
SO4 mg/l 6.942 68.71 0.0706
NO3 mg/l 0.0307 0.295 0.0014
F mg/l 0.018 0.173 0.0008
Br mg/l 0.541 5.268 0.0149
SiO2 mg/l 0.0068 0.067 9.39E-05
B(boron) mg/l 1.223 7.242 0.553
TDS mg/l 280.8 2737 7.51
Conductivity @25C μS 581.5 5069 18.94
pH pH 9.50 9.545 8.77
Osmotic pressure bar 0.229/0.22 2.182/2.09 0.007/0.01
(DS1/Pitzer)
LSI/SDSI 0.70/0.81
(continued)
Annexes 731

Concentrate Pass Product Pass


Feed to Pass 2 2 2
Ions Units 40 26 20
CaSO4/SrSO4% % 0.1%/0.1%
BaSO4/SiO2% % 2.3%/0.0%

Pass/stage data Units Pass 1/Stage 1


Lead element type TM820V-400
Last element type TM820V-400
Total elements 9544
Total vessels 1193
Elements per vessel 8
Feed flow m3/day 246,911
Product flow m3/day 111,110
Average flux l/m2/h 13.005
Brine flow m3/day 135,801
Recovery % % 45.00%
Feed pressure bar 50.72
dP elements bar 0.603
Brine pressure bar 50.12
Permeate pressure bar 1.0
Feed TDS mg/l 34,559
Perm TDS mg/l 270.0
Lead element
Feed flow m3/day 207.0
Product flow m3/day 22.27
Product TDS mg/l 109.3
Flux l/m2/h 24.87
Last element
Product flow m3/day 3.982
Product TDS mg/l 944.6
Brine/product ratio – 28.59
Brine flow m3/day 113.8
Net driving pressure bar 3.446
Beta – 1.038

Pass/stage data Pass 2 Stage 1 Stage 2 Stage 3


Lead element type TM720L-440
Last element type TM720L-440
Total elements 2910 1740 840 330
Total vessels 485 290 140 55
Elements per vessel 6
(continued)
732 5 Reverse Osmosis Membrane System: Core Process of SWRO

Pass/stage data Pass 2 Stage 1 Stage 2 Stage 3


Units
Feed flow m3/day 111,110 43,594 18,007
Product flow m3/day 67,519 25,587 6884
Average flux l/m2/h 39.55 31.05 21.26
Brine flow m3/day 43,594 18,007 11,122
Recovery % % 60.77% 58.69% 38.23%
Feed pressure bar 11.739 10.030 8.704
dP elements bar 1.709 1.326 1.674
Brine pressure bar 10.030 8.704 7.029
Permeate pressure bar 1.000 1.000 1.000
Feed TDS mg/l 269.5 679.6 1628
Perm TDS mg/l 3.426 10.966 31.32
Lead element Stage 1 Stage 2 Stage 3
Feed flow m3/day 383.1 311.4 327.4
Product flow m3/day 42.52 34.27 25.31
Product TDS mg/l 2.333 7.038 22.47
Flux l/m2/h 43.35 34.93 25.80
Last element Stage 1 Stage 2 Stage 3
Product flow m3/day 35.50 26.64 16.658
Product TDS mg/l 5.718 18.803 48.09
Brine/product ratio – 4.235 4.827 12.140
Brine flow m3/day 150.3 128.6 202.2
Net driving pressure bar 8.501 6.343 3.938
Beta – 1.182 1.156 1.066

References
1. Matsuura T., Sourirajan S., Theory and Practice of Reverse Osmosis - Part 1, IDA Interna-
tional Desalination Association, 1985.
2. Loeb S., “The Loeb-Sourirajan Membrane : How it Came About,” in Synthetic Membranes
Volume 1 Desalination, American Chemical Society, 1981, pp. 1-9 Chapter 1 .
3. Sourirajan S., Reverse Osmosis, Logos Press Limited, 1971.
4. Gulf General Atomic Incorporated, Westmoreland J.C., “US Patent 3,367,594 - SPIRALLY
WRAPPED REVERSE OSMOSIS MEMBRANE CELL,” 1968.
5. Gulf General Atomic Incorporated, Merten U., “US Patent 3,386,583 - REVERSE OSMOSIS
MEMBRANE MODULE,” 1968.
6. Gulf General Atomic Incorporated, Bray D. T., “US Patent 3,417,870 - REVERSE OSMOSIS
PURIFICATION APPARATUS,” 1968.
7. FilmTec Corporation, Cadotte J.E., “US - Patent 4,277,344 INTERFACIALLY
SYNTHESIZED REVERSE OSMOSIS MEMBRANE,” 1981.
8. Large Diameter Element Membrane Consortium - Bartels C., Shelby I.- Hydranautics; Peery
M., Hallan M. - FilmTec; Metcalfe P. - Toray; Knappe P.- Trisep , Henthorne J - CH2M Hill,
“Desalination and Water Purification Research and Development Report No. 114 - Industry
Consortium Analysis of Large Reverse Osmosis/Nanofiltration Element Diameters,” U.-
S. Department of the Interior - Bureau of Reclamation, 2005.
References 733

9. Jeong B.H., Hoek E.M.V., Yan Y., Subramani A.,Huang X., Hurwitz G., Gosh A.K., Jawor
A., “Interfacial polymerization of thin film nanocomposites: A new concept for reverse
osmosis membranes,” Journal of Membrane Science, vol. 294, pp. 1 - 7, 2007.
10. NanoH2O, Inc, Kurth C.J., Koehler J.A., Zhou M., Holmberg B.A., Burk R.L., “US Patent
8,177,978 B2 - REVERSE OSMOSIS MEMBRANES,” 2012.
11. Kurth C.J., Burk R.L., Green J., “Improving Seawater Desalination With Nanocomposite
Membranes,” IDA Journal, pp. 26 - 31, Third Quater 2010.
12. Misdan N., Lau W.J., Ismail A.F., “Seawater Reverse Osmosis (SWRO) desalination by thin-
film composite membrane - Current development, challenges and future prospects,” Desalina-
tion, vol. 287, pp. 228-237, 2012.
13. Ludwig, H., Reverse Osmosis Seawater Desalination Volume 2, Springer, 2022
14. Saline Water Conversion Corporation SWCC - Mayan Kutty P.C., Dalvi A.G.I., Al Jerfaley T.
H., Thankachan T.S., “Evolution of Bromine in Sea Water MSF Desalination Plants - Tech.
Report No. SWCC (RDC)-10,” SWWC, 1991.
15. Shemer H., Semiat R., “Impact of halogen based disinfectants in seawater on polyamide RO
membranes,” Desalination, vol. 273, no. 1, pp. 179-183, 2011.
16. Bousher A., Brimblecombe P., Midgley D., “Rate of Hypobromite Formation in Chlorinated
Seawater,” vol. 20, no. 7, pp. 865-870, 1986.
17. Glater J., Hong S., Elimelech M., “The search for a chlorine-resistant reverse osmosis
membrane,” Desalination, vol. 95, pp. 325 - 345, 1994.
18. Do V.T., Tang C.Y., Reinhard M. Leckie J.O., “Effects of Chlorine Exposure Conditions on
Physicochemical Properties and Performance of Polyamide Membrane - Mechanisms and
Implications,” Environmental Science and Technology, vol. 46, pp. 13184 - 13912, 2012.
19. Sandin R., Ferrero E., Repolles C., Navea S., Bacardit J., Espinos J. P., Malfeito J.J., “Reverse
osmosis membranes oxidation by hypochlorite and chlorine dioxide: spectroscopic
techniques vs. Fujiwara test,” Desalination and Water Treatment, vol. 51, pp. 318 - 327, 2013.
20. Ettori A., Gaudichet-Maurin E., Aimar P., Causserand C., “Pilot scale study of chlorination-
induced transport property changes of a seawater reverse osmosis membrane,” Desalination,
vol. 311, pp. 24 - 30, 2013.
21. Kwon Y., Joksimovic R., Kim I., Leckie J.O., “Effect of bromide on the chlorination of a
polyamide membrane,” Desalination, vol. 280, no. 1 - 3, pp. 80 - 86, 2011.
22. McCray S.B., Glater J., Cutchan J.W., “The Effect of pH and Halogens on the Stability of
Reverse Osmosis Membranes, UCLA-ENG-8115,” Water Resources Center Report, 1981.
23. Adams W.R., “The Effects of Chlorine Dioxide on Reverse Osmosis Membranes,” Desalina-
tion, vol. 78, pp. 439 - 453, 1990.
24. Eriksson P.K, “Chlorine Dioxide for Prevention of Biofouling of Polyamide RO Elements,” in
IDA World Congress, Maspalomas, Spain, 2007.
25. Tseng T.J., Cheng R.C., Tanuwidjaja D., Wattier K.L., “Evaluating ClO2 in Seawater
Desalination Pretreatment,” IDA Journal, pp. 54 - 60, Second Quater 2010.
26. Eriksson P.K., Dimotsis G.L., “Field Experience: Can Chlorine Dioxide be Used sd a Biocide
in RO Plants ?,” IDA Journal, pp. 14 - 20, First Quarter 2012.
27. Ferrero E., Llansana A., Ayala V., Malfeito J.J, “Chlorite and Chlorate Effect on the Reverse
Osmosis Membranes Performance,” in IDA World Congress, Maspalomas, Gran Canaria,
Spain, 2007.
28. Applegate L.E., Erkenbrecher C.W., Winters H., “New chloramine process to control
aftergrowth and biofouling in permasep B-10 RO surface seawater plants,” vol. 74, pp.
51-67, 1989.
29. E.I. Du Pont de Nemours and Company, Applegate L.E., Erkenbrecher C.W., Winters H., “US
Patent 4,988,444 - PREVENTION OF BIOFOULING OF REVERSE OSMOSIS
MEMBRANES,” 1991.
30. Maugin T., Croue J.P.,Marinas B., “RO Desalination of Chloraminated Sea and Brackish
Water: Membrane Performance,” IDA World Congress - Perth, Western Australia, 2011.
31. Gabelich C.J., Frankin J.C., Gerringer F.W., Ishida K.P., Suffet I.H., “Enhanced oxidation of
polyamide membranes using monochloramine and ferrous iron,” Journal of Membrane Sci-
ence, vol. 258, pp. 64 - 70, 2005.
734 5 Reverse Osmosis Membrane System: Core Process of SWRO

32. Cran M.J., Bigger S.W., Gray S.R., “Degradation of polyamide reverse osmosis membranes in
the presence of chloramine,” Desalination, vol. 283, pp. 58 - 63, 2011.
33. Hydranautics, “Technical Application Bulletin 115 - Potential Use of ClO2 as a Disinfectant
for Polyamide RO/NF Membranes,” October 2013.
34. Mc Cray S.B., Vilker V.L., Nobe K., “Reverse osmosis cellulose acetate membranes
II. Dependence of transport properties on acetyl content,” vol. 59, pp. 317 - 330, 1991.
35. Vos K.D., Burris F.O., Riley R.L., “Kinetic Study of the Hydrolysis of Cellulose Acetate in the
pH Range of 2 - 10,” Journal of Applied Polymer Science, vol. 10, pp. 825 - 832, 1966.
36. Kumano A., Matsui Y., Numata K., Fujiwara N., Iwahashi H., Nagai M., “Performance change
formula of cellulose triacetate hollow fiber RO membranes due to oxidation and hydrolysis,”
Desalination , vol. 96, pp. 451 - 457, 1994.
37. Tanaka S., Kuzumoto H., Sekino M., “Using Chlorine Dioxide for Trihalomethane Control on
RO Seawater Desalination System,” in IDA World Congress on Desalination and Water
Sciences Vol III Session 6, Abu Dhabi, 1995.
38. Fujiwara N., Numuta K., Kumano A., Ogino Y., Nagai M., Iwahashi H., “The effect of heavy
metal ions on the oxidation of cellulose triacetate membranes,” Desalination, vol. 96, pp. 431 -
439, 1994.
39. Fujiwara N., Matsuyama H., “Optimization of the intermittent chlorine injection (ICI) method
for seawater desalination RO plants,” Desalination, vol. 229, pp. 231 - 244, 2008.
40. Sommariva C., Attenborough A., Poggi F., Al-Mahdi A.M., Hori T., Tokunaga T.K., Ito Y.,
Maeda Y., “Matching Hollow- Fiber With Spiral-Wound Membranes: Process Compatibility
and Optimization,” IDA Journal, vol. I, no. Second Quarter, pp. 40 - 44, 2012.
41. Lonsdale H.K., Merten U., Riley R.L., “Transport properties of cellulose acetate osmotic
membranes,” Journal of Applied Polymer Science, vol. 9, no. 4, pp. 1341 - 1362, 1965.
42. Sherwood T.K., Brian P.L.T., Fisher R.E., “Desalination by Reverse Osmosis,” Industrial &
Engineering Chemistry Fundamentals, vol. 6, no. 1, pp. 2 - 12, 1967.
43. Kedem O., Katchalsky A., “Thermodynamic analysis of the permeability of biological
membranes for non-electrolytes,” Biochimica et Biophysica Acta, vol. 27, pp. 229 - 246, 1958.
44. Spiegler K.S., Kedem O., “Thermodynamics of hyperfiltration (reverse osmosisi): criteria for
efficient membranes,” Desalination, vol. 1, no. 4, pp. 311 - 326, 1966.
45. ASTM International, “D4516 - 00 Standard Practice for Standardizing Reverse Osmosis
Performance Data,” 2010.
46. Sekino M., “Study of Membrane Transport for Memodiafilter and Reverse Osmosis Module,”
vol. 2, pp. 125-133, 2013.
47. Taniguchi M., Kimura S., “ Estimation of Transport Parameters of RO Membranes for
Seawater Desalination,” AIChE Journal, vol. 47, no. 10, pp. 1967-1973, 2000.
48. Schock G., Miquel A, “Mass transfer and pressure loss in spiral wound modules,” Desalina-
tion, vol. 64, p. 339–352, 1987.
49. Poisson A., Papaud A., “Diffusion coefficients of major ions in seawater,” Marine Chemistry,
vol. 13, pp. 265 - 280, 1983.
50. Saltonstall C.W., Lawrence R.W., “Calculation of the expected performance of reverse
osmosis plants,” Desalination, Bd. 42, pp. 247 - 253, 1982.
51. Uemura T., Henmi M., “Thin-Film Composite Membranes for Reverse Osmosis,” in
Advanced Membrane Technology and Applications – Part 1 – Membranes and Applications
in Water and Wastewater, John Wiley & Sons, 2008, pp. 3 - 19.
52. Wilf M., The Guidebook to Membrane Desalination Technology, Balaban Desalination
Publications, 2007.
53. Dow Filmtec, “Tech Fact Form-Nr. 609-00436-1211 – How to Calculate Your Bromide
Rejection Levels”.
54. Mesmer R.E., Baes C.F., Sweeton F.H., “Acidity Measurements at Elevated Temperature.
VI. Boric Acid Equilibria,” Inorganic Chemistry, vol. 11, no. 3, pp. 537 - 543, 1972.
55. Kim J., Hyung H., Wilf M .,Park J., Brown J., “Boron Rejection by Reverse Osmosis
Membranes: National Reconnaissance and Mechanism Study– Desalination and Water Purifi-
cation Research and Development Program Report No. 127,” U.S Department of the Interior -
Bureau of Reclamation - Technical Service Center, Denver,Colorado, 2009.
References 735

56. Hyung H., Kim J., “A mechanistic study on boron rejection by sea water reverse osmosis
membranes,” Journal of Membrane Science, vol. 286, pp. 269 - 278, 2006.
57. Franks R., Neculau M., Garrote R., Bartels C., Egea R.J., Carrion M.L., Saura A.J.T., Prieto
M.R., “Analyzing three years of SWRO operation at elevated feed pH to save energy and
improve boron rejection,” in The International Desalination Association World Congress on
Desalination and Water Reuse, Tianjin, China, 2013.
58. Chillon Arias M.F., Bru L.V., Rico D.P., Galvan P.V., “Kinetic behaviour of sodium and
boron in brackish water membranes,” Journal of Membrane Science, vol. 368, p. 86 –
94, 2011.
59. Busch M., Mickols W.E., Jons S., Redondo J., Witte J.D., “Boron removal in sea water
desalination,” in International Desalination Association- World Congress on Desalination
and Water Reuse, Paradise Island, Bahamas, 2003.
60. Tomioka H., Taniguchi M., Okazaki M., Gotoh S., Tadahiro U., Kurihara M., “Milestone of
High Boron Rejection Seawater Membrane,” in IDA World Congress on Desalination and
Water Reuse, Singapore, 2005.
61. Henmi M., Tomioka H., Kawakami T., “Performance Advancement of High Boron Removal
Seawater RO Membranes,” in IDA World Congress of Desalination and Water Reuse,
Maspalomas,Gran Canaria, Spain, 2007.
62. Bartels C. R., Rybar S., Andes K., Franks R., “Optimized Removal of Boron and Other
Specific Contaminants by SWRO Membranes,” in IDA World Congress on Desalination and
Water Reuse , Dubai, United Arab Emirates, 2009.
63. Farhat A., Ahmad F., Hital N., Arafat H. A., “Boron removal in new generation reverse
osmosis (RO) membranes using two-pass RO without pH adjustment,” Desalination, vol.
310, pp. 50 - 59, 2013.
64. Ludwig H., “Reverse osmosis in the desalination of brackish water and sea water,” Desalina-
tion, vol. 36, pp. 153 - 178, 1981.
65. Dow FILMTEC, “FILMTEC Reverse Osmosis Membranes,” in Technical Manual - Section 3
- System Design, Dow Water & Process Solutions, pp. 70 - 100.
66. Mickols W.E., Busch M., Maeda Y., Tonner Y., “A Novel Design Approach for Seawater
Plants,” in IDA World Congress on Desalination and Water Reuse, Singapore, 2005.
67. Penate B., Garcia-Rodriguez L., “Reverse osmosis hybrid membrane inter-stage design: A
comparative performance assessment,” Desalination, vol. 281, pp. 354 - 363, 2011.
68. Desalination Systems, Inc., Bray D.T, “US Patent 4,046,685 - SIMULTANEOUS PRODUC-
TION OF MULTIPLE GRADES OF PURIFIED WATER BY REVERSE OSMOSIS,” 1977.
69. Rybar S., Boda R., Bartels C., “Split partial second pass design for SWRO plants,” Desalina-
tion and Water Treatment, vol. 13, pp. 186 - 194, 2010.
70. Kurihara M., Yamamura H., Nakanishi T., “High recovery/high pressure membranes for brine
conversion SWRO process development and its performance data,” Desalination, vol. 125, pp.
9-15, 1999.
71. Kurihara M., Yamamura H., Nakanishi T., Jinno S., “Operation and reliability of very high-
recovery seawater technologies by brine conversion two-stage RO desalination system,”
Desalination, vol. 138, pp. 191-199, 2001.
72. Moch I., “The Case For And Feasibility Of Very High Recovery Sea Water Reverse Osmosis
Plants,” Desalination & Water Reuse Quaterly, vol. 10/3, pp. 44-52, 2000.
73. Gründisch A., Schneider B.P., “Optimising energy consumption in SWRO systems with brine
concentrators,” Desalination, vol. 138, pp. 223-229, 2001.
74. von Gottberg A., Pang A., Talavera J.L., “Optimizing Water Recovery and Energy Consump-
tion for Seawater RO Systems,” GE Water & Process Technologies - Technical Paper
TP1021EN 0510.
75. Toray Industries, Inc., Yamamura H., Kurihara M., Maeda K., “US Patent 6,187,200 B1 -
APPARATUS AND METHOD FOR MULTISTAGE REVERSE OSMOSIS
SEPARATION,” 2001.
736 5 Reverse Osmosis Membrane System: Core Process of SWRO

76. Toray Industries, Inc., Yamamura H., Kurihara M., Maeda K., “EUROPEAN PATENT
SPECIFICATION EP 1 161 981 B1 - Method for multistage reverse osmosisi
separation,” 2008.
77. MEDRC Series of R & D Reports - MEDRC Project: 03-AS-001- Fane A.G., “REVIEW OF
COLLOIDAL FOULING IN SPIRAL WOUND MODULES,” The Middle East Desalination
Research Center, Muscat, Sultanate of Oman, 2008.
78. Li H., Fane A.G., Coster H.G.L., Vigneswaran S., “An Assessment of Depolarisation Models
of Crossflow Microfiltration by Direct Observation through the Membrane,” Journal of
Membrane Science, vol. 172, pp. 135 - 147, 2000.
79. Field R.W., Wu D., Howell J.A., Gupta B.B., “Critical flux concept for microfiltration
fouling,” Journal of Membrane Science, vol. 100, pp. 259 - 272, 1995.
80. Bacchin P., Aimar P., Field R.,, “Critical and sustainable fluxes: Theory, experiments and
applications,” Journal of Membrane Science, vol. 281, pp. 42-69, 2006.
81. Neal P.R., Li H., Fane A.G., Wiley D.F., “The effect of filament orientation on critical flux and
particle deposition in spacer-filled channels,” Journal of Membrane Science , vol. 214, pp.
165-178, 2003.
82. Schippers J.C., Hanemaayer J.H., Smolders C.A., Kostense A., “PREDICTING FLUX
DECLINE OF REVERSE OSMOSIS MEMBRANES,” Desalination , vol. 38, pp. 339 -
348, 1981.
83. MEDRC Series of R & D Reports - MEDRC Project: 97-AS-004b - Winters H., “Identification
of Critical Flux and Cross Flow Conditions for the Control of Bacterial and Organic Fouling of
Seawater Reverse Osmosis Membranes,” The Middle East Desalination Research Center,
Muscat, Sultanate of Oman, 2001.
84. Winters H., “USE OF CRITICAL FLUX VALUES TO CONTROL MICROBIAL FOULING
IN REVERSE OSMOSIS DESALINATION,” in NACE International, CORROSION 2000 ,
Orlando, Florida, 2000.
85. Chong T.H., Wong F.S., Fane A.G., “Implications of critical flux and cake enhanced osmotic
pressure (CEOP) on colloidal fouling in reverse osmosisi: Experimental observations,” vol.
314, pp. 101 - 111, 2008.
86. Zhang Y.P, Chong T.H., Fane A.G., Law A., Coster H.G.L., Winters H., “Implications of
enhancing critical flux of particulates by AC fields in RO desalination and reclamation,”
Desalination, Bd. 220, pp. 371 - 379, 2008.
87. Li H., Fane A.G., Coster H.G.L., Vigneswaran S., “Observation of deposition and removal
behaviour of submicron bacteria on the membrane surface during cross flow microfiltration,”
Journal of Membrane Science , vol. 217, pp. 29 - 41, 2003.
88. Vrouwenvelder J.S., van Paasen J.A.M., van Agtmaal J.M.C., van Loosdrecht M.C.M.,
Kruithof J.C., “A critical flux to avoid biofouling of spiral wound nanofiltration and reverse
osmosisi membranes: Fact or fiction ?,” Journal of Membrane science , vol. 326, pp. 36 -
44, 2009.
89. Hoek E.M.V., Kim A.S., Elimelech M., “Influence of Crossflow Membrane Filter Geometry
and Shear Rate on Colloidal Fouling in Reverse Osmosis and Nanofiltration Separations,”
Environmental Engineering Science, vol. 19, no. 6, pp. 357 - 372, 2002.
90. Hoek E.M.V., Elimelech M., “Cake-Enhanced Concentration Polarization: A New Fouling
Mechanism for Salt- Rejecting Membranes,” Environmental Science & Technology, vol.
37, pp. 5581 - 5588, 2003.
91. Sutzkofer I., Hasson D., Semiat R., “Simple technique for measuring the concentration
polarization level in a reverse osmosis system,” Desalination, vol. 131, pp. 117 - 127, 2000.
92. Chong T.H., Wong F.S., Fane A.G., “Enhanced concentration polarization by unstirred
fouling layers in reverse osmosis: Detection by sodium chloride tracer response technique,”
Journal of Membrane Science, vol. 287, pp. 198 - 210, 2007.
93. Chong T.H., Reverse Osmosis Desalination and Reclamation - Control of Colloidal Fouling
and Biofouling - PhD - Thesis, Singapore: Nanyang Technological University - School of
Civil and Environmental Engineering , 2008.
References 737

94. Ju Y., Hong S., “Nano-colloidal fouling mechanism in seawater reverse osmosis process
evaluated by cake resistance simulator-modified fouling index nanofiltration,” Desalination,
vol. 343, pp. 88 - 96, 2014.
95. ASTM International, “Annual Book of ASTM Standards, Section 11- Water and Environmen-
tal Technology,” no. Volume11.01 Water (I), Volume 11.02 Water (II).
96. Alhadidi A.. Kemperman A.J.B.,Blankert B., Schippers J.C., Wessling M., van der Meer W.G.
J., “Silt Density Index and Modified Fouling Index relation, and effect of pressure, temperature
and membrane resistance,” Desalination, vol. 273, pp. 48-56, 2011.
97. Alhadidi A., Blankert B., Kemperman A.J.B., Schurer R., Schippers J.C., Wessling M., van
der Meer W.G.J., “Limitations, improvements and alternatives of the silt density index,”
Desalination and Water Treatment, vol. 51, pp. 1104 - 1113, 2013.
98. Mosset A., Bonnelye V., Petry M., Sanz M.A., “The sensitivity of SDI analysis: from RO feed
water to raw water,” Desalination, vol. 222, pp. 17 - 23, 2008.
99. Garcia M, Sanz J., Carulla C., Nebot E., Ortega J.M., Casanas A., Lubian L.M., Quevedo N.,
“ASSESSMENT OF MFI AS SEAWATER PRETREATMENT DESIGN TOOL AT
CARBONERAS DEMONSTRATION PLANT,” in IDA World Congress, Perth, 2011.
100. Gasia-Bruch E., Busch M., Salinas-Rodriguez S.G., Kennedy M.D., “Improving Fouling
Indices Measurements and Modeling Their Relevance,” IDA Journal, vol. I Fourth Quarter
2011, pp. 58 - 73, 2011.
101. Boerlage S.F.E., Kennedy M.D., Dickson M.R., El-Hodali D.E.Y., Schippers J.C., "The
modified fouling index using ultrafiltration membranes (MFI-UF): characterisation, filtration
mechanism and proposed reference membrane," Journal of Membrane Science , vol. 197, pp.
1-21, 2002.
102. Boerlage S.F.E., Kennedy M., Aniye M.P., Schippers J.C., “Applications of the MFI-UF to
measure and predict particulate fouling in RO systems,” Journal of Membrane Science, vol.
220, pp. 97-116, 2003.
103. Salinas-Rodriguez S.G, “PARTICULATE AND ORGANIC MATTER FOULING OF SEA-
WATER REVERSE OSMOSIS SYSTEMS - Dissertation,” CRC Press/Balkema, Delft Uni-
versity of Technology/UNESCO - IHE, 2011.
104. Boerlage S. F. E., Kennedy M., Tarawneh Z., De Faber R., Schippers J.C., “Development of
the MFI-UF in constant flux filtration,” Desalination, vol. 161, pp. 103 - 113, 2004.
105. Salinas-Rodriguez S.G., Kennedy M.D., Amy G.L., Schippers J.C., “Flux dependency of
particulate/colloidal fouling in seawater reverse osmosis systems,” Desalination and Water
Treatment, vol. 42, pp. 155-162, 2012.
106. Jeong S., Vigneswaran S., Leiknes T.O., “A NEW APPROACH TO ESTIMATE THE
BIOFOULING POTENTIAL IN SEAWATER DESALINATION PLANT USING
MODIFIED FOULING INDEX,” The International Desalination Association World Con-
gress- San Diego USA, 2015.
107. MEDRC Series of R&D Reports - MEDRC Project : 07-AS-004 -Althuluth M., “FURTHER
DEVELOPMENT OF THE MODIFIED FOULING INDEX (MFI-UF) AT CONSTANT
FLUX FOR SWRO APPLICATIONS,” The Middle East Desalination Research Center,
Muscat, Sultanate of Oman, 2009.
108. Jaques D.F, Bourland B.I. , “A Study of Solubility of Strontium Sulfate,” Society of Petroleum
Engineers Journal, vol. 23, no. 2, pp. 292 - 300, 1983.
109. Ezuber H.M., “Prediction of Strontium Sulfate Scale Formation in Oilfield Environment,”
Journal of ASTM International, vol. 4, no. 6, 2007.
110. Ezuber H.M. , “Prediction of Barium Sulfate Scale Formation in Oilfield Environment,”
Journal of ASTM International, vol. 8, no. 2, 2009.
111. DuPont, “Bulletin 5050,” in Permasep Products Engineering Manual, 1992, pp. Page
7, Figure 4.
112. Dow Water & Process Solutions - Dow FILMTEC, “Section 2.4 Scaling calculations - 2.4.6
Calcium Fluoride Scale Prevention - Figure 2.9,” in FILMTEC Reverse Osmosis Membranes -
Technical Manual, p. 44 .
738 5 Reverse Osmosis Membrane System: Core Process of SWRO

113. Garand A., Mucci A., “The solubility of fluorite as a function of ionic strength and solution
composition at 25  C and 1 atm total pressure,” Marine chemistry, vol. 91, pp. 27 - 35, 2004.
114. McGee K.A., Hostetler P.B., “Activity - Product Constants of Brucite from 10 to 90  C,”
Journal of Research of the U.S. Geological Survey, vol. 5, no. 2, pp. 227 - 233, 1977.
115. Stiff H.A., Davis L.E., “A Method for Predicting the Tendency of Oil Field Waters to Deposit
Calcium Carbonate,” Petroleum Transactions, AIME, vol. 195, pp. 213 - 216, 1952.
116. Millero F.J., Huang F.,Graham T., Pierrot D, “The dissociation of carbonic acid in NaCl
solutions as a function of concentration and temperature,” Geochimica et Cosmochimica Acta,
vol. 71, pp. 46 - 54, 2007.
117. Plummer N.L., Busenberg E., “The solubilities of calcite, aragonite and vaterite in CO2-H2O
solutions between 0 and 90 C, and an evaluation of the aqueous model for the system CaCO3-
CO2-H2O,” Geochimics et Cosmochimica Acta, vol. 46, pp. 1011 - 1040, 1982.
118. Harned H.S., Owen B.B., The Physical Chemistry of Electrolytic Solutions, 3rd edition,
New York: Reinhold Publishing Corp., 1958.
119. Langelier W.F., “The Analytical Control of Anti-Corrosion Water Treatment,” Journal Amer-
ican Water Works Association, vol. 28, pp. 1500 - 1521, 1936.
120. Larson T.E., Buswell A.M., “Calcium Carbonate Saturation Index and Alkalinity
Interpretations,” Journal of the American Water Works Association, vol. 34, no. 11, pp.
1667 - 1684, 1942.
121. Waly T.K.A, Minimizing the use of chemicals to control scaling in SWRO: Improved
prediction of the scaling potential of Calcium Carbonate, Dissertation at Delft University of
Technology, UNESCO-IHE Institute for Water Education: CRC Press/Balkema, 2011.
122. He S., Kan A.T., Tomson M.B, Inhibition of Mineral Scale Precipitation by Polymers in
Amjad Z.(ed.), Water Soluble Polymers - Solution Properties and Applications,
Springer, 1998.
123. He S., Kan A.T., Tomson M.B, “Inhibition of calcium carbonate precipitation in NaCl brines
from 25 to 90  C,” Applied Geochemistry, vol. 14, pp. 17 - 25, 1999.
124. Boerlage S.F.E, Kennedy M.D., Bremere I., Witkamp G.I., van der Hoek J.P., Schippers J.C.,
“Stable barium sulphate supersaturation in reverse osmosis,” Journal of Membrane Science ,
vol. 179, pp. 53 - 68, 2000.
125. Boerlage S.F.E, Kennedy M.D., Bremere I., Witkamp G.I., van der Hoek J.P., Schippers J.C,
“The scaling potential of barium sulphate in reverse osmosis systems,” Journal of Membrane
Science, vol. 197, pp. 251 - 268, 2002.
126. Kan, A., Fu G., Fan C., Tomson M., “Quantitative evaluation of calcium sulfate precipitation
kinetics in the presence and absence of scale inhibitors,” in SPE International Symposium on
Oilfield Chemistry, 2009.
127. Waly T., Kennedy M.D., Witkamp G., Amy G., Schippers J.C., “The role of inorganic ions in
the calcium carbonate scaling of seawater reverse osmosis systems,” Desalination, vol.
284, pp. 279-287, 2012.
128. Sheikholeslami R., Ong H.W.K., “Kinetics and thermodynamics of calcium carbonate and
calcium sulfate at salinities up to 1.5 M,” Desalination, vol. 157, pp. 217 - 234, 2003.
129. Boerlage S.F.E, Kennedy M.D., Witkamp G.I., van der Hoek J.P., Schippers J.C, “BaSO4
solubility prediction in reverse osmosis membrane systems,” Journal of Membrane Science,
vol. 159, pp. 47 - 59, 1999.
130. Waly T., Kennedy M.D., Witkamp G-J., Amy G., Schippers J.C., “Predicting and measure-
ment of pH of seawater reverse osmosis concentrates,” Desalination, vol. 280, pp. 27 -
32, 2011.
131. Behrends H., Baumgarten S., Matz B., Schill J., Stoffel B., “Technical Features of High-
pressure Pumps for RO Facilities,” in Internation Desalination Association IDA - World
Congress on Desalination and Water Reuse, Singapore, 2005.
132. Stoffel B., Assessing the Energy Efficiency of Pumps and Pump Units, Elsevier,
Europump, 2015.
References 739

133. de la Torre A., “Efficiency optimization in SWRO plant: high efficiency & low maintenence
pumps,” Desalination , vol. 221, pp. 151 - 157, 2008.
134. Sulzer Pumps Ltd., Centrifugal Pump Handbook, Elsevier, 2010.
135. Moch I., Oklejas M., Terrasi K., Oklejas R.A., “Advanced High Efficiency Energy Recovery,”
in IDA World Congress on Desalination and Water Reuse, Singapore, 2005.
136. Sanchez J.M.S., Castillo N.S., Castillo R.S, “Mathematical Model for Isobaric Energy Recov-
ery Devices,” in IDA World Congress, Maspalomas, Gran Canaria, Spain, 2007.
137. Energy Recovery Inc., “ERI Technical Bulletin - Isobaric Device Mixing - Doc. No. 80088-
01,” 2008.
138. Stover R.L., Andrews B., “Isobaric Energy-Recovery Devices : Past, Present , and Future,”
IDA Journal, vol. I, no. First Quartal 2012, pp. 38 - 43, 2012.
139. Kochanowski W., Schwarz G., Klemm T., Bross S., “Mixing and Overflushing - Important
Considerations on Using Pressure Exchanger Technology,” in IDA World congress,
Maspalomas, Gran Canaria-Spain, 2007.
140. Oklejas E., Leachman L.M., Kitzmiller R.T., Seisan A., Kadaj E., “A Novel Equipment
Centralization Schema Reduces the Cost of Permeate,” in IDA World Congress, Maspalomas,
Gran Canaria - Spain, 2007.
141. Hermant B.M., Basu O.D, “Comparison of Reaction Rates and Relative Efficiencies for
Various Dechlorination Chemicals,” Journal of Environmental Engineering, vol. 139, no.
4, pp. 522 - 529, 2013.
142. Dow Water & Process Solutions - Dow FILMTEC, “Section 2.6.3 Chlorination/Dechlorina-
tion,” in FILMTEC Reverse Osmosis Membranes - Technical Manual, 2015, p. 60.
143. Saeed M.O., “Effect of dechlorination point location and residual chlorine on biofouling in a
seawater reverse osmosis plant,” Desalination, vol. 143, pp. 229 - 235, 2002.
144. Hydranautics, “Technical Application Bulletin 110 - Chlorination in RO Seawater Supply
Lines and Pretreatment Processes,” 2014.
145. Chesters S.P., Armstrong M.W., Fazel M., Wilson R., Golding D.A., “RO Membrane
Cleaning, Past, Present,Future - Innovations for Improving RO Plant Operation Efficiency,”
in The International Desalination Association World Congress on Desalination and Water
Reuse 2013, Tianjin, China, 2013.
146. Chesters S.P., Pena N., Gallego S., Fazel M., Armstrong M.W., del Vigo F., “Results from
99 Seawater RO Membrane Autopsies,” in IDA World Congress, Perth, Western
Australia, 2011.
147. Wilson R., Fazel M., Jarrige S., Chesters S., “Air Bubbles Enhance Membrane Cleaning: A
Future Perspective,” in The Internationale Desalination Association World Congress on
Desalination and Water Reuse 2013, Tianjin, China, 2013.
148. Partlan E., “Dissolved Carbon Dioxide for Scale Removal in Reverse Osmosis - Thesis at the
Graduate School of Clemson University,” Clemson, South Carolina, USA, 2013.
149. Rietman B.M, “Cleaning spiral wound membrane moduls with a two phase solution - Thesis at
the Delft University of Technology,” Delft, Netherlands, 2013.
150. Qin J., Liberman B., Kekre K.A., “Direct Osmosis for Reverse Osmosis Fouling Control:
Principles, Applications and Recent Developments,” The Open Chemical Engineering Jour-
nal, vol. 3, pp. 8 - 16, 2009.
151. I. Liberman, “RO MEMBRANE CLEANING METHOD”. United States Patent US 7,658,852
B2, 9 February 2010.
152. Hydranautics, “Technical Service Bulletin TSB 125 - Reverse Direction Cleaning of RO
Elements,” 2013.
153. Liberman B., Faigon M., Hefer D., “Flexible 3-centre setup saves costs in Ashkelon,”
Desalination and Water Reuse, vol. 14/4, pp. 19 - 21, 2005.

You might also like