Sie sind auf Seite 1von 157

Institut für Thermodynamik

und Wärmetechnik

Universität Stuttgart

Convective Heat Transfer Fouling of


Aqueous Solutions on
Modified Surfaces

Abdullah K. O. Al-Janabi
ISBN 978-3-00-035657-5
Convective Heat Transfer Fouling of
Aqueous Solutions on
Modified Surfaces

A thesis accepted by the Faculty of Energy Technology, Process Engineering and Biological
Engineering of the University of Stuttgart in Partial Fulfilment of the Requirements
for the Degree of
Doctor of Engineering Sciences (Dr.-Ing.)

By

Abdullah K. O. Al-Janabi
born in Babylon, Iraq

Chairman: Prof. Dr.-Ing. Eberhard Göde


Main referee: Prof. Dr. Dr.-Ing. habil. H. Müller-Steinhagen
Co-referee: Prof. Dr. techn. Günter Scheffknecht

Date of oral examination: 4th April 2011

Institute of Thermodynamics and Thermal Engineering


University of Stuttgart, Germany
2011
Gedruckt mit Unterstützung des Deutschen Akademischen Austauschdienstes
Dedication

I dedicate this thesis to my:

Family
for providing unlimited love,
support and enthusiasm throughout my life

Wife
for her love, encouragement and
for being the joy of my life

Beloved country, Iraq

i
ii
Abstract

ABSTRACT

The present research study was part of the European project "MEDESOL" entitled
“Seawater desalination by innovative solar-powered membrane-distillation system”. The
project aimed at developing a stand-alone desalination unit to produce fresh water with a
maximum of 50 m3/day. Several components such as suitable membrane and efficient solar
collectors had to be developed as well as a plate heat exchanger for a maximum life
expectancy with least deposition occurrence on its surfaces. The contribution of ITW to this
project was mainly concerned with the development of innovative anti-fouling heat transfer
surfaces that could substantially i) reduce the design and operation penalties, and ii) increase
the membrane distillation efficiency. To achieve this goal, this study endeavoured to address
some outstanding questions such as:

1) The impact of surface texture in terms of roughness and structured grooves on the
fouling propensity.
2) The effect of various modified surfaces on adhesion mechanisms between the
deposits and the heat transfer surfaces.
3) A criterion that would predict whether a modified surface would foul or not.

Accordingly, a comprehensive and rigorous set of fouling runs was performed with
calcium sulphate solution as working fluid. The operating conditions followed closely to
those expected in the membrane distillation desalination unit i.e. a fluid velocity of up to 0.3
m/s and a maximum surface temperature of 90oC.
The experimental results showed that i) increasing surface roughness causes a
significant reduction of the nucleation correction factor ( φ ). As a consequence, a higher
fouling rate would be expected due to the reduction in the energy barrier limit of
crystallization; ii) in case of grooved surfaces, the generated eddies within the crossed
grooves have vortexes which are opposite to those of the mainstream. This would overcome
the adhesion forces between the crystals and the heat transfer surfaces and gave rise to longer
induction time and lower fouling rate; and iii) the deposition process was strongly affected by
altering the surface energy properties, particularly the electron donor component ( γ 2− ) for the

iii
Abstract

coatings investigated in this study. Lower initial fouling rate and longer induction time were
the main features of coatings having higher γ 2− .
A new criterion has also been developed that predicts whether a modified surface will
foul or not. In this criterion, the total interaction energy depends strongly on the Lewis acid-
base energy while the contribution of the Lifshitz-van der Waals energy is marginal. A
validation of the new criterion was carried out by comparison with the experimental results of
the present study as well as some previous investigations of precipitation and biofouling. The
validation results confirmed the reliability and accuracy of the new criterion.
Finally, the coatings that performed best in terms of fouling mitigation in the
laboratory tests were selected for field assessment in two successive phases. No sign of
fouling was observed in either phase. However, in phase 1, substantial flaking-off of the
coatings occurred in particular around the inlet and outlet sections of the heat exchanger as
well as at the contact points between plates. In phase 2, a thinner coating was applied with a
more roughened surface which resulted in far better stickiness of coatings to the surface.

iv
Kurzfassung

KURZFASSUNG

Die Forschungsarbeit "Seawater desalination by innovative solar-powered membrane-


distillation system" war Teil des europäischen Projektes MEDESOL. Ziel dieser Arbeit war
die Entwicklung einer autark operierenden Entsalzungsanlage für die Frischwasserproduktion
von maximal 50 m3/Tag. Verschiedene Komponenten wie zum Beispiel adäquate
Membranen und effiziente Solarkollektoren mussten entwickelt werden. Ebenso wurde ein
verbesserter Plattenwärmeübertrager entwickelt, um eine maximale Lebensdauer mit
minimalem Ablagerungsvorkommen auf den Flächen zu gewährleisten.
Die Mitwirkung des ITW an diesem Projekt betraf hauptsächlich die Entwicklung
innovativer Anti-fouling-Wärmeübertragungsflächen, die im Wesentlichen i) die Mehrkosten
aufgrund der konstruktiven Ausführung und der Betriebsbedingungen reduzieren können,
sowie ii) die Effizienz der Membrandistillation steigern sollen.
Um dieses Ziel zu erreichen, befasst sich diese Arbeit mit folgenden Punkten:

1) Einfluss der Oberflächenbeschaffenheit, d.h. Rauheit und strukturierte Riefen, auf


die Neigung zur Ablagerung.
2) Auswirkung diverser modifizierter Oberflächen auf die Adhäsion zwischen den
Ablagerungen und der Wärmeübertragungsfläche.
3) Aufstellen eines Kriteriums, mit dem bestimmt werden kann, ob sich auf einer
veränderten Oberfläche Ablagerungen ansammeln.

Es wurden eine Reihe von Versuchen mit einer Calcium-Sulphat-Lösung als


Betriebsmittel durchgeführt. Die Betriebsbedingungen, d.h. eine Strömungsgeschwindigkeit
von bis zu 0.3 m/s und eine maximale Oberflächentemperatur von 90˚C, waren ähnlich wie
die in der Membran-Destillation-Entsalzungsanlage.
Die Versuchsergebnisse haben gezeigt, dass
i) die Zunahme der Oberflächenrauhigkeit eine signifikante Abnahme des
Keimbildungs-Korrektur-Faktors ( φ ) verursacht. Als Folge davon wird erwartet,
dass die Foulingrate aufgrund der Senkung des Energieniveaus zur Kristallbildung
zunimmt,

v
Kurzfassung

ii) die an den Querrillen erzeugten Wirbel sorgen dafür, dass die Adhäsionskräfte
zwischen den Kristallen und der Oberfläche überwunden werden,
iii) der Ablagerungsprozess von den sich ändernden Eigenschaften der
Oberflächenergie stark beeinflusst wurde. In dieser Arbeit wurde vor allem die
Elektronengeberkomponente ( γ 2− ) der Beschichtungen untersucht. Eine
anfänglich niedrigere Foulingrate und eine längere Induktionszeit waren die
Hauptmerkmale der Beschichtungen, die einen höheren γ 2− -Wert aufweisen.

Ebenso wurde ein neues Kriterium entwickelt, welches vorhersagt, ob sich auf einer
modifizierten Oberfläche Ablagerungen ansammeln. Dabei hängt die
Wechselwirkungsenergie sehr stark von der Lewis Säure-Basis-Energie ab, während die
Wirkung der Lifshitz-van der Waals-Energie gering ist. Die Zuverlässigkeit und die
Genauigkeit des neuen Kriteriums wurden durch die experimentellen Ergebnisse sowohl in
dieser Arbeit als auch in früheren Arbeiten bestätigt.
Schließlich wurden die zwei besten Beschichtungen für Feldversuche ausgesucht und
in einem Plattenwärmeübertrager in zwei aufeinanderfolgenden Phasen eingesetzt. Obwohl in
beiden Phasen keine Spur von Ablagerungen zu beobachten war, löste sich in der ersten
Phase, im Bereich des Ein-und Ausgangs des Wärmeübertragers und an den Kontaktpunkten
zwischen den Platten, die Beschichtung ab. In der zweiten Phase wurde jedoch eine dünnere
Beschichtung auf eine rauere Oberfläche angebracht, was zu einer besseren Haftung führte.

vi
Acknowledgements

ACKNOWLEDGEMENTS

I gratefully acknowledge the help and support of the following people without whom I would
have unable to do this work.

First and foremost, Particular thanks are due to my supervisor, Professor H. Müller-
Steinhagen, for his invaluable comments, guidance and support throughout this research
project. Without his input, none of this work would have been possible.

Special thanks goes to Dr. M.R. Malayeri, my co-supervisor and the head of heat exchanger
fouling and cleaning research group at the Institute of Thermodynamics and Thermal
Engineering (ITW), University of Stuttgart. Special thanks for his guidance, discussions
whenever I needed, weekly meetings, and the encouragement throughout my PhD.

I am also grateful to Dr. Q. Zhao of The University of Dundee for answering my questions
that I had about surface energy.

The help and warm hospitality of all members of staff and technicians at ITW has been very
much appreciated.

I would like to thank sincerely DAAD (Deutscher Akademischer Austauschdienst) for the PhD
research studentship. I thank also the European Commission for the financial support of the
MEDESOL Project (European project, Contract no. 036986) through which part of my
research has been financed.

Last but not least, I am especially thankful for the love and support of my family and friends for
making my stay aboard possible.

vii
Nomenclature

NOMENCLATURE

A Hamaker constant [J]


As heat transfer area [m2]
B bias limit [-]
B1 constant [J/m2]
B2 constant [m4K/W2 s]
Co constant [m2/N2]
C constant [m2K/W]
c concentration [g/L]
d depth [m]
D diameter [m]
E constant [N/m2]
Ea activation energy [J/mol]
Eea electronegativity [kJ/mol]
ΔE132
AB
Lewis acid-base interaction energy [J]

ΔE Br Brownian energy [J]

ΔE132
EL
electrostatic double layer interaction energy [J]

ΔE132
LW
Lifshitz-van der Waals interaction energy [J]

ΔE132
TOT
total interaction energy [J]

ΔE131
TOT
total cohesion energy [J]

ΔGcrit critical free energy [J]

ΔGcrit
het
heterogeneous critical free energy [J]

ΔGcrit
hom
homogenous critical free energy [J]
ΔGs surface excess free energy [J]
ΔGv volume excess free energy [J]
ΔG f interfacial energy [J]

H separation distance between the interacting bodies [m]


Ho minimum equilibrium distance between a particle and a surface [m]

viii
Nomenclature

I ionic strength [mol/L]


J het heterogeneous nucleation rate [m-3 s-1]
K constant [--]
k thermal conductivity [W/m K]
ka adhesion coefficient [m/s]
kf fluid thermal conductivity [W/m K]
kB Boltzmann constant [J/K]
Kf surface texture parameter [m-2]
kr reaction rate constant [m4/kg.s]
krem proportional constant for removal [m3/kg.s]
L length [m]
M number of independent variables [-]
N number of observations [-]
P precision limit [-]
q& heat flux [W/m2]
rcrit critical radius of nucleus [m]
R particle radius [m]
Ra roughness profile [m]
Rf fouling resistance [m2 K/W]
R *f asymptotic fouling resistance [m2 K/W]

Rg universal gas constant [J/K mol]


Rq root mean square roughness of a rough surface [m]
Rw wall resistance [m2 K/W]
Rz mean roughness depth [m]
S supersaturation ratio [-]
t time [min.]
T temperature [oC]
tind induction time [min.]
U oveall heat transfer coefficient [W/m2K]
U uncertainty [-]
v velocity [m/s]
w width [m]

ix
Nomenclature

W work of adhesion [J]


x distance [m]
X mean value of a set of N observations [-]

Y profile hight [m]


z ion's valence [-]

Dimensionless Number

Nu Nusselt number [–]


Pr Prandtl number [–]
Re Reynolds number [–]

Greek symbols

α heat transfer coefficient [W/m2K]


β mass transfer coefficient [m2/s]
γ surface energy [J/m2]
γ+ electron acceptor component of surface energy [J/m2]
γ- electron donor component of surface energy [J/m2]
γ12 solid/liquid interface surface energy [J/m2]
γ2 solid/vapour interface surface energy [J/m2]
γ3 liquid/vapour interface surface tension [J/m2]
γAB Lewis acid-base surface energy [J/m2]
γLW Lifshitz-van der Waals surface energy [J/m2]
δ linear expansion coefficient of the deposit layer [K-1]
δ+ Lewis acid polarity [-]
δ− Lewis base polarity [-]
ε electrical permittivity of the solution [C/V.m]
ζ zeta potential [mV]
θ contact angle [o]
κ −1 Debye length [m]
λ correlation length pertaining to water molecules [m]
λ12 spreading coefficient [J/m2]
μ chemical potential [kJ/mol]

x
Nomenclature

ν volume of nucleus [m3]


ξ adhesive strength [N/m2]
σf shear strength of the deposit layer [N/m2]

τ shear stress [N/m2]


φ nucleation correction factor [-]
Λ surface energy parameter [-]

Subscripts

b bulk
f foulant
i interface
o initial
s surface
t time
th thermocouple
d deposition
r removal
c clean
ss stainless steel
co coated surface

Superscript
AB Lewis acid/base
d dispersive
LW Lifshitz-van der Waals
p polar
het heterogeneous
hom homogenous

Abbreviations
AISI American Iron and Steel Institute
Al aluminum

xi
Nomenclature

BN boron nitride
Ca 2+ calcium ions

Ca(NO3)2·4 H2O calcium nitrate tetrahydrate


CaCl2 calcium chloride
CaCO3 calcium carbonate
CaSO4 calcium sulphate anhydrite
CaSO4.1/2H2O calcium sulphate hemihydrate
CaSO4.2H2O calcium sulphate dihydrate
CCD Charge Coupled Device
CFU colony-forming unit
CH2I2 diiodomethane
C2H6O2 ethylene glycol
Cr chrome
CrN chrome nitride
Cu copper
DIN German Institute for Standardization
DLC Diamond Like Carbon
DSA Drop Shape Analysis
EDTA Ethylene-Diamine-Tetra-acetic Acid
EPDM Ethylene Propylene Diene M-class rubber
F fluorine
FEP Fluorinated Ethylene Propylene
FEPA Federation of European Producers of Abrasives
H hydrogen
HCL hydrochloric acid
H2O water
ITW Institute for Thermodynamics and Thermal Engineering
Li2SO4 lithium sulphate
MD Membrane Distillation
Mg(OH)2 magnesium hydroxide
MgSO4 magnesium sulphate
MoS 22+ molybdenum disulfide ion

xii
Nomenclature

Ms(I) brass alloy


Ms(II) brass alloy
Na +2 sodium ions
Na2SO4 sodium sulphate
NaCl sodium chloride
NaNO3 sodium nitrate
Ni nickel
Ni-P nickel-phosphor
NO 3+ nitrate ions
O oxygen
P phosphor
PECVD Plasma Enhanced Chemical Vapour Deposition
PFA Perfluoroalkoxy
PHE plate heat exchanger
PID proportional integral derivative controller
PTFE Polytetrafluoro Ethylene
SB solvent based coated surface
SEM Scanning Electron Microscope
Si silicon
SICAN Si-DLC
SICON Si-O-DLC
SS stainless steel
TiN titanium nitride
TiO2 titanium dioxide
WB water based coated surface

xiii
Table of Contents

TABLE OF CONTENTS

1. INTRODUCTION 1
1.1 Short Description of the MEDESOL Project 1
1.2 Heat Exchanger Fouling 3
1.2.1 Fouling mechanisms 5
1.2.2 Crystallization fouling on heat transfer surfaces 6
1.2.3 Mitigation techniques 8
1.3 Research Objectives 9
1.4 Scope of Present Study 9

2. LITERATURE REVIEW 11
2.1 Surface Modification Approaches to Mitigate Fouling 11
2.1.1 Surface geometry modifications 11
2.1.1.1 Roughened surfaces 12
2.1.1.2 Structured surfaces 13
2.1.2 Surface treatment through alteration of surface energy 14
2.1.2.1 Surface energy concept 14
2.1.2.2 Intermolecular interaction energies 19
2.1.2.3 Effect of different surface properties on fouling mechanisms 23
2.1.2.4 Fouling models with consideration of surface energy properties 27

3. EXPERIMENTAL SET-UP AND PROCEDURE 33


3.1 Experimental Apparatus 33
3.1.1 Test rig 33
3.1.2 Heat transfer specimen 35
3.1.3 Data acquisition and reduction 36
3.2 Surface Characterizations 37
3.2.1 Surface roughness 37
3.2.2 Contact angle 38
3.2.3 Surface energy 39
3.3 Experimental Procedure 40
3.3.1 Test solution and preparation procedure 40

xiv
Table of Contents

3.3.2 Clean and fouling heat transfer coefficients 42


3.3.3 Determination of basic fouling parameters 42
3.3.4 Error analysis 44
3.3.5 Matrix of operating conditions 45

4. EXPERIMENTAL RESULTS AND DISCUSSION 47


4.1 Alteration of Surface Geometry 47
4.1.1 Fouling of roughened surfaces 47
4.1.1.1 Investigation of various roughened surfaces 47
4.1.1.2 Effect of contact angle on fouling mechanisms 50
4.1.2 Fouling of structured surface geometries 56
4.1.2.1 Heat transfer over grooved surfaces 57
4.1.2.2 Clean heat transfer coefficient 58
4.1.2.3 Fouling of grooved surfaces 60
4.1.2.4 Effect of groove dimensions on fouling mechanisms 62
4.2 Alteration of the Surface Energy Related Properties 64
4.2.1 Surface characterizations 64
4.2.1.1 Coating composition and preparation 64
4.2.1.2 Physical surface properties 65
4.2.1.3 Resistance of coatings to abrasion and corrosion 68
4.2.2 Fouling investigation of modified surfaces 69
4.2.2.1 Solvent/water based coated surfaces 69
4.2.2.2 Electroless Ni-P coated surfaces 70
4.2.2.3 Nano-structured coated surfaces 71
4.2.3 Post-fouling analysis of the deposit layer on modified surfaces 75
4.2.4 Influence of surface energy on fouling propensity 79
4.2.5 Reproducibility of fouling propensity 86

5. THEORETICAL STUDY 88
5.1 Validation of the Previous Surface Energy Fouling Models 88
5.2 A New Simplified Surface Energy Fouling Criterion 90
5.3 Validation of the New Surface Energy Fouling Criterion 95

6. FIELD INVESTIGATION OF COATED SURFACES 102

xv
Table of Contents

6.1 Heat Exchanger Specifications 102


6.2 Examination of the Modified Surfaces in the MEDESOL-I Prototype 103

7. CONCLUSIONS AND FUTURE WORK 109


7.1 Conclusions 109
7.1.1 Alteration of the surface geometry 109
7.1.2 Alteration of the surface energy related properties 110
7.2 Future Work 113

LIST OF REFERENCES 114


APPENDIXES 125
LIST OF PUBLICATIONS 131

xvi
1 Introduction

Chapter 1
INTRODUCTION

1.1 Short Description of the MEDESOL Project


The present study was part of the European project entitled “seawater desalination by
innovative solar-powered membrane-distillation system (MEDESOL)”. The project aimed at
developing an environmentally friendly and innovative concept of multistage Membrane
Distillation (MD) in order to minimize specific thermal energy input and membrane area
required. The required heat has to be provided by solar thermal collectors. Another objective
was to discern the technical feasibility of producing 0.5 to 50 m3/day potable water from
seawater by integrating several membrane distillation modules. A conceptual schematic
diagram of the MEDESOL is presented in Figure 1.1. The prototype consists of three main
components, i.e. solar collector (1), heat exchanger (2) and the membrane (3).

1 B
90oC
2
3
cold side
hot side

C
45oC
distilled water

A
25oC 4
saline water

1 solar collector
2 plate heat exchanger
3 membrane distillation
4 distilled water tank

Fig. 1.1 Conceptual description of the MEDESOL prototype

Membrane distillation (MD) is a desalination technique where saline water is heated


to increase its vapour pressure, which in turn generates the difference between the partial
pressure at both sides of the membrane. Hot water evaporates through non-wetted pores of

1
1 Introduction

hydrophobic membranes. The passing vapour is then condensed on a cooler surface to


produce fresh water. The overall functionality of the process is illustrated in Figure 1.1. The
seawater (A) at low temperature of (e.g. 25oC) is heated with the hot water coming from the
solar collector to provide warm seawater (B, e.g. 90oC) to be fed into the evaporator channel
of the membrane. The partial pressure difference caused by the temperature difference across
both sides of the membrane is the driving force for the steam (fresh water) passing through
the membrane. The seawater is cooled through the evaporator channel arriving to (C, e.g.
45oC). The efficiency of the MD module increases as the inlet temperature of seawater goes
up. Thus, the unit should permit seawater to heat up as close as possible to its saturation
temperature without the risk of deposit formation. For the sake of compactness and maximum
life expectancy, it was decided to use a plate and frame heat exchanger (PHE) which has the
following advantages over conventional shell & tube heat exchangers:
i) typically only one-fourth to one-tenth of the space of that of shell & tube heat
exchangers is required (Carlson, 1992);
ii) it can be easily disassembled, inspected, maintained, cleaned if necessary and
reassembled again; and
iii) high rate of heat transfer is expected due to high turbulence and low thermal
resistance of the plates (Carlson, 1992).
These advantages have increased the market shape of plate and frame heat
exchangers. Nevertheless, the formation of deposit on heat transfer surfaces would still
reduce the performance of the plate heat exchanger. Figure 1.2 shows typically the formation
of deposits on the surface of a plate heat exchanger.

Fig. 1.2 Deposit build-up of CaSO4 in a plate heat exchanger (Bansal et al., 2000)

2
1 Introduction

The formation of deposits on heat transfer surfaces is the least understood problem in
the design of heat exchangers due to large number of influential parameters with poorly
understood interaction. Within the framework of the MEDESOL project, the present study
aimed at addressing the following objectives concerning the formation of deposit on heat
transfer surfaces:

1. The impact of surface finish and texture on fouling propensity


2. Surface characteristics that can influence fouling propensity most
3. Available and suitable coatings technique for the plate heat exchanger
4. Thermal and mechanical stabilities of the modified surfaces for long-term
operation of the heat exchanger

Answers to these objectives would provide technical guidelines for future research in
this field through modification of i) surface geometries, and ii) surface energy related
properties.

1.2 Heat Exchanger Fouling


Fouling of heat exchangers is a common industrial problem which may occur in
various processes such as oil refineries, desalination units, food/dairy industries and power
plants. It may be any type of undesired deposits such as particulate matter, rust, organic and
inorganic deposits that form on heat transfer surfaces during the life time of the heat
exchanger. Fouling of heat exchangers can lead to i) increase in thermal resistance to heat
transmission due to the lower thermal conductivity of the deposit layer which would range
from 0.5 to 3.0 W/m.K (Bott, 1995), ii) increase in pressure drop across the heat exchanger.
Deposits can be tenacious “hard scale” that may be difficult to remove, or soft deposits that
are sometimes called “sludge”.
To characterize the thermal performance of heat transfer surfaces which are subject to
fouling, the overall heat transfer coefficient for the two heat exchanging fluids has been
defined according to Eq. (1.1):

1 ⎛ 1 ⎞A 1
= ⎜⎜ + R f ,1 ⎟⎟ s 2 + Rw + + R f ,2 (1.1)
U ⎝ α1 ⎠ As1 α2

3
1 Introduction

Here, U, α, As, Rf and Rwall are overall heat transfer coefficient, film heat transfer
coefficient, heat transfer areas, fouling resistances and thermal resistance of the separating
wall, respectively. If fouling occurs on one side only and if all operating parameters of the
heat exchanging fluids are maintained constant, then the fouling resistance may be
determined by

1 1
Rf = − (1.2)
Ut Uo

In Eq. (1.2), subscripts “t” and “o” denote the overall heat transfer coefficient at any
time and at the beginning of the experiment/operation, when the heat transfer surface is
considered to be clean. Figure 1.3 depicts a typical variation of fouling resistance versus time
for a stainless steel surface and CaSO4 as foulant (Malayeri et al., 2009). Three specific
features may be identified in this graph, i.e. the induction period that elapses before a
substantial increase in fouling resistance takes place and the fouling rate which is the slope of
the curve where the fouling resistance increases continuously during the fouling period.

-3
1x10
R *f asymptotic fouling resistance
-4
8x10
fouling resistance [m K/W]

-4
2

6x10
Induction period

-4
4x10
slop = initial fouling rate q& = 80 kW/m2
-4
2x10 v = 0.15 m/s
cb= 3.75 g/L
Fouling period o
Tb= 40 C
0 CaSO4
AISI 304 BA
-4
-2x10
0 500 1000 1500 2000 2500

time [min.]

Fig. 1.3 Induction and fouling periods in a typical fouling resistant curve
(Malayeri et al., 2009)

4
1 Introduction

1.2.1 Fouling mechanisms


Fouling can form due to various mechanisms based on the chemical/physical
conditions under which the deposit layer builds up on the surface and can be classified into
the following mechanisms:

1. Crystallization fouling: It occurs when salts precipitate on a heat transfer surface, only if
the surface temperature is sufficient to cause supersaturation of the salt.
2. Particulate fouling: The accumulation of solid particles suspended in the process stream
onto the heat transfer surface results in particulate fouling. In some cases, the deposition
occurs due to gravity, in which case it is referred to as sedimentation fouling.
3. Chemical reaction at the heat transfer surface may yield to deposition as in the case of
polymerization and cracking.
4. Corrosion fouling occurs when the surface material reacts with the process fluid or
beneath the already formed deposit which in turn gives rise to corrosion product. It can be
limited or prevented altogether by the correct choice of heat transfer surface materials.
5. Biofouling is the attachment and growth of organisms to heat transfer surfaces. The
organisms may be divided into micro-organisms such as bacteria and macro-organisms
such as algae and seaweed.

As the prototype plant of the MEDESOL project contained a filter which is able to
remove suspended particles thus only mineral salts, e.g. CaSO4, CaCO3 and Mg(OH)2 are
expected to deposit on the surface. Patel and Finan (1999) showed that two types of scale
could form in desalination processes depending on the operating conditions:

• Alkaline soft scale, e.g. CaCO3 and Mg(OH)2


• Non-alkaline typical hard scale such as CaSO4

Fouling due to calcium carbonate can be controlled by varying the acidity of the water
(Kazmierczat et al., 1982), while calcium sulphate precipitation is not significantly affected
by the pH of the water. Accordingly, in the present study, calcium sulphate aqueous solution
has been considered as the working fluid for simulating the seawater in the experimental
study. It is therefore indispensable to understand the mechanism of CaSO4 crystallization
fouling as well as the influential parameters.

5
1 Introduction

1.2.2 Crystallization fouling on heat transfer surfaces


The formation of crystal deposits on heat transfer surfaces is a common phenomenon
where aqueous solutions are involved. The deposition of salts from solution is influenced by
two processes of nucleation and crystal growth which can only occur when the solution is
supersaturated at its respective temperature. There are two ways by which a process of
aqueous solution in a heat transfer equipment may become supersaturated; i) when a solution
containing a dissolved salt of normal solubility such as NaCl and NaNO3 is cooled to a level
below its solubility and ii) a dissolved salt of inverse solubility such as CaSO4, CaCO3 and
Li2SO4 is heated up. The latter is the main cause of supersaturation conditions in most heat
exchangers. However, being in supersaturation state alone is not sufficient for crystallization.
A number of nuclei or seeds that act as centers of crystallization must exist on the surface.
These seeds may either occur spontaneously or be induced artificially. In industrial processes,
both conditions are likely to affect the formation of deposits on heat transfer surfaces.
In general, crystallographers divide crystal nucleation into two main categories as
shown in Figure 1.4, i.e. homogeneous and heterogeneous nucleation. Both of them take
place in the absence of solution-own crystals and they are collectively known as primary
nucleation. In industrial processes, nevertheless, nucleation occurs even at a very low
supersaturation when solution-own crystals exist, a process which is referred as the secondary
nucleation. It may be caused by fragments of broken crystals or the presence of small
particles in the solution which may act as new nuclei.

homogeneous
nucleation
crystal growth

primary
superstauration

nucleation
nucleation

heterogeneous
nucleation
secondary
nucleation

Fig. 1.4 Crystal formation steps (Mullin, 1993).

As time goes on the seeds formed during the process increase in size until they
become visible. This process is known as crystal growth which can occur through a series of
consecutive steps: bulk diffusion, surface adsorption, surface reaction, and finally integration
of ions or molecules into the crystal lattice (Mullin, 1993). Generally, the mechanisms of

6
1 Introduction

crystal growth can be broadly discussed as two-step process; bulk diffusion, in which the
crystal growth is governed by the concentration difference between the bulk solution and the
solid-liquid interface. It is quite possible that mass transfer is the rate-determining step.

⎛ dm ⎞
⎜ ⎟ = β (cb − ci ) (1.3)
⎝ dt ⎠ d

The integration of ions through the boundary layer into the crystal lattice is
considered to be the second step of crystal growth. Mathematically this step can be
considered as a chemical reaction. The rate at which ions are integrated in the crystal lattice,
can be given by (Bott, 1995):

⎛ dm ⎞
⎟ = k r (ci − c s )
n
⎜ (1.4)
⎝ dt ⎠ d

where k r and c are the reaction rate constant (m4/kg.s) and the concentration (g/L),
respectively. While the subscripts ''i'' and ''s'' denote conditions at solid-liquid interface and at
saturation, respectively. For calcium sulphate, "n" has been determined to be equal to 2 (Liu
and Nancollas, 1970).
Removal is a simultaneous process to deposition thus it is quite possible that some of
the deposits will be swept away by shear exerted by the fluid. Stronger shear forces will not
entirely prevent the deposits from forming but will lead to thinner and firmer deposits
(Kukulka and Devgun, 2007). The removal term depends on velocity, deposit thickness and
strength of the deposit layer as:

⎛ dm ⎞ τ
⎜ ⎟ = k rem δ (Bohnet, 1987) (1.5)
⎝ dt ⎠ r σf

where τ , σ f and δ are the shear stress (N/m2), shear strength of the deposit layer (N/m2) and

the linear expansion coefficient of the deposit layer (K-1), respectively (Bohnet, 1987). The
rate of deposit formation on a surface can then be considered as difference between the rate
of solid formation and that of solid removal as shown in Figure 1.5.

7
1 Introduction

Operating conditions, e.g. Physicochemical properties, e.g.


• surface and bulk temperatures • type of salt
• flow velocity • supersaturation

deposition process removal process

Surface
characterizations
• surface roughness
• surface texture
• surface energy
heat transfer surface

Fig. 1.5 hypothetical sketch of the deposit formation

Generally, it can be said that crystallization fouling is strongly dependent upon the
following parameters:
• Foulant physicochemical properties.
• Operating conditions such as flow velocity, bulk and surface temperatures.
• Surface characteristics which refer to the substrate material and surface texture
where the deposit is attached.

1.2.3 Mitigation techniques


Numerous mechanical, chemical and physical approaches have been developed for
mitigating or controlling mineral scale formation including sponge balls, enhanced heat
exchanger surfaces and chemical inhibitors such as dispersing and chelating agents (Panchal
and Knudsen, 1998). In addition, there have been studies to utilize magnetic and electric
fields, ultrasound, vortex flows and sudden pressure changes to mitigate fouling. However,
the use of such techniques has not given consistent results (Cho et al., 2003).
In recent years technological advancements in surface treatment have given new
impetus for developing innovative coating technologies and materials for a wide range of
fouling problems under various modes of heat transfer such as convection and boiling
(Müller-Steinhagen and Zhao, 1997; Zhao et al., 2005a). Accordingly, there is increasing
attention to utilize modified surfaces which are characterized by low surface energy, weak
adhesive or even non-sticking properties. In these approaches, the notion is to decrease the

8
1 Introduction

adhesive strength between deposits and heat transfer surfaces as the deposit formation on a
heat transfer surface always results from the interaction between the deposits and the solid
surface. Widespread industrial application to-date is limited to the use of high-grade
metals/alloys or polymer coatings to mitigate corrosion and to some extent fouling.

1.3 Research Objectives


The main objectives of the present study are to
1. minimize foulant adhesion energies by altering the surface geometry and energy
related properties of the heat transfer surfaces followed by evaluating the
contribution of various intermolecular interaction energies to the adhesion
process,
2. provide environmentally-friendly non-stickable modified surfaces for utilization
in the MEDESOL project with maximum life-expectancy and free of the need for
regular cleaning,
3. develop a fouling criterion that identifies the surface energy component which is
responsible for the adhesion process.
To achieve these objectives, firstly, a comprehensive set of experiments has been
carried out to include i) the investigation of the effect of surface geometry in terms of
roughened and grooved surfaces on crystallization fouling mechanisms, ii) the modification
of the surface energy related properties of AISI 304 BA stainless steel surface by using
different coating techniques and iii) field assessment of the best performing coatings in
laboratory tests in the MEDESOL prototype. Secondly, a simultaneous theoretical study to
identify the most influential component of surface energy which dominates the adhesion
process as well as to develop a new qualitative criterion to indicate whether a modified
surface would foul or not.

1.4 Scope of Present Study


From the industrial application point of view, this study contributes to a better
understanding of how the surface characteristics in terms of geometry and energy properties
affect the mechanisms of precipitation fouling. This has been fulfilled both experimentally
and theoretically through 6 different work-packages (WP) as shown in Figure 1.6. Each
work-package is divided into several tasks.

9
1 Introduction

In this study, Chapter 2 presents a literature survey on i) the principle of surface


energy and ii) state-of-the-art of fouling research with respect to the impact of surface
geometry and surface energy properties on crystallization fouling. Chapter 3 describes the
experimental set up followed by a detailed elucidation of surfaces investigated in this study.
The experimental procedure to conduct fouling runs is also presented in this chapter.
Furthermore, the definitions used to describe and evaluate the fouling behaviour are outlined
as well as a description of how the experimental uncertainty is calculated. Chapter 4
describes fouling experimental results as a function of surface properties and operating
conditions. Furthermore information such as deposit layer, deposit morphology and
reproducibility of the fouling runs for various modified surfaces are discussed. The
theoretical study is presented in Chapter 5 for evaluating the contribution of the interfacial
energies to the total interaction energy. Chapter 6 includes details about the examination of
the modified surfaces in the MEDESOL prototype. Finally, the main conclusions are
summarized in Chapter 7 followed by suggested future work.

Literature Survey

Research Methodology

surface geometry surface energy


modification modification

Pre-fouling surface characterization

Convective heat transfer fouling tests

Post-fouling analysis of fouled surfaces

Analysis of results & Theoretical study

Field investigation of coated surfaces in the


MEDESOL prototype desalination unit

Fig. 1.6 Interrelation of the work-packages of this study

10
2 Literature Review

Chapter 2
LITERATURE REVIEW

In the past few years, several experimental investigations demonstrated the significant
effects of surface energy properties and geometry on the reduction of mainly crystallization
fouling. In most, but certainly not all, cases lower surface energy resulted in reduced
propensity of the surface to foul. While there have been some successful attempts at fouling
mitigation with modified surfaces, many problems remain unanswered such as
irreproducibility of results and degradation of the modified surface performance. Better
understanding of the interfacial energies between deposits and surfaces could provide the
required insights into the adhesion mechanisms on modified surfaces as well as to serve as a
guide to create conditions that will reduce the attractive energy and consequently prevent
fouling. This chapter aims to discern important parameters in terms of geometrical and
surface energy properties that may influence the deposition process on various surfaces.

2.1 Surface Modification Approaches to Mitigate Fouling


When a particle such as a crystal or bacterium reaches the surface, first it contacts the
surface and if the adhesion energies are strong enough then will stick to the surface. The
interfacial properties, particularly surface roughness, topography, wettability and surface
energy of a solid surface, are the most important parameters which play a decisive role in the
adhesion process (Wu, 1982). It is therefore indispensable to get insight into various available
modification techniques that influence these parameters.

2.1.1 Surface geometry modifications


For enhanced heat transfer purposes the alteration of surface geometry is perhaps the
simplest and yet highly effective technique which results in increased contact surface area as
well as stronger agitation of the boundary layer. Examples include artificial surface
roughness (irregularities) in terms of elementary cavities on the bare surface or secondary
effects due to erosion of the surface and/or the periodic disturbance promoters such as fins or
grooves. Such modifications can substantially affect the fouling characteristics as a result of
their direct impact on mechanical interlocking adhesion.

11
2 Literature Review

2.1.1.1 Roughened surfaces


Surface roughness is defined as closely spaced random irregularities of a surface as
shown in Figure 2.1. Sometimes it is considered an undesirable property as it may cause
friction and drag. It can also have a contrary impact and be beneficial as it may allow surfaces
to increase convective heat and mass transfer due to the agitation of the viscous sub-layer
(Rahman and Gui, 1993). However, whether transfer rates are increased or decreased depends
largely on the nature of the roughness, i.e. distribution of the roughness elements, shape, size
and orientation (Mahato and Shemilt, 1968).

peak

pit

Fig. 2.1 Surface roughness irregularities

Undoubtedly, under fouling conditions increasing surface roughness, in most cases,


causes a significant increase in fouling rate during the fouling period (Gunn, 1980; Kazi et
al., 2001; Förster and Bohnet, 2001). The effect of roughness on fouling can be explained as:
i) a large surface roughness increases the level of turbulence and consequently
enhances the mass transport which in turn leads to provide nucleation sites that encourage the
formation of the initial deposit. Mott (1991) concluded that smooth surfaces (glass and
electropolished AISI 316 stainless steel) had 35% less deposits than untreated AISI 316
stainless steel surface. Kazi et al. (2001) similarly concluded that the higher surface
roughness, the more severe fouling would be due to more available sites for nucleation. In
terms of induction time, it has also been found that reducing the mean roughness depth of
stainless steel surface enlarges the induction time of the CaSO4 crystallization process
(Zettler et al., 2005; Geddert et al., 2007). In addition, irregularities of a roughened surface
can increase the contact surface area, i.e. alteration of surface wettability which in turn can
encourage the heterogeneous nucleation (Strickland-Constable, 1968).
ii) a consequence of the enhanced nucleation on a rough surface is the formation of a
stronger adhesive bond (Yoon and Lund, 1994). This is due to the fact that a rough surface
has a greater effective surface energy compared to smooth ones (Geddert et al., 2007). On the

12
2 Literature Review

other hand, stronger adhesion should occur on the rough surface since it would undergo a
greater decrease in effective surface energy. Keysar et al. (1994) showed that the adhesion
strength of a calcite fouling layer on the metal surface (AISI 1020) was remarkably
influenced by the degree of surface roughness. The tensile stress required to disbond the
calcite deposits adhering to a roughened metal surface was found to be as much as 30 times
higher than that required to disbond the same deposits adhering to a smooth surface.
Despite such reasonings, there are still several uncertainties that should be addressed
through more rigorous experimental work such as, i) the lack of systematic attempts to
discern the impact of surface roughness on the adhesion mechanisms of calcium sulphate and
ii) lack of experimental work to see whether roughness would have similar impact on
modified surfaces as for the untreated stainless steel surface.

2.1.1.2 Structured surfaces


Flow interruption created in flow at spatial intervals is a common practice for
enhancing heat transfer in heat exchangers. Fins or grooves are well-known examples of
spatial intervals geometry surfaces that can serve as disturbance promoters by increasing fluid
mixing and extending heated surface area. Thus, such techniques are frequently exploited to
manufacture more compact and smaller heat exchangers with higher efficiency. Over the past
decades, tremendous efforts have been devoted to better understanding of flow mixing and
heat transfer enhancement in channels with different geometries, like grooved, wavy and
corrugated channels (Ghaddar et al., 1986; Greiner, 1991; Wang and Vanka, 1995). While
excellent performance is often reported for compact heat exchangers under clean conditions,
opposite trends were observed when subjected to fouling, since the narrow channels are
readily plugged by deposits. For instance, Heinzel et al. (2007) found that most of the
particulate deposition occurs on the edges of the rectangular channels as well as at the inlet
grid wall of micro heat exchangers. In addition, Benzinger et al. (2005) studied the
precipitation of calcium carbonate within micro-channels under the influence of ultrasound as
means of deposition removal. They showed that the operating time of micro heat exchangers
without or with less deposit could be extended using ultrasound. Having said that, although
several attempts have been made to investigate particulate fouling of grooved surfaces, there
is virtually no such result on crystallization fouling. Furthermore, no fundamental study has
been undertaken so far to investigate how the surface geometry of structured grooves (size,
shape and direction to the flow) would influence crystallization fouling.

13
2 Literature Review

2.1.2 Surface treatment through alteration of surface energy


2.1.2.1 Surface energy concept
Surface energy defines as the excess energy at the surface of a material compared to
the bulk. The molecules within a material have less energy due to their strong interactions
with like molecules in all directions as shown in Figure 2.2, while the molecules at the
surface are attracted less and they will try to reduce this free energy by interacting with
molecules in an adjacent phase.
There is no direct method to determine the surface energy of solid surfaces other than
through the measurements of contact angle combined with an appropriate theoretical
approach. It is therefore indispensable to understand the basis of the contact angle and the
theoretical approaches used to determine the surface energy.

Fig. 2.2 Diagram of forces on molecules at the top of a surface and in the bulk material

In a pioneering work, Young (1805) proposed an equation to describe the force


balance between a solid (1), a liquid (2) and its vapour (3) as shown in Figure 2.3:

γ 2 = γ 12 + γ 3 cos θ (2.1)

γ3

vapour

liquid θ γ2

γ12
solid

Fig. 2.3 Contact angle due to interfacial free energies at the boundaries between solid, liquid
and vapour

14
2 Literature Review

where γ 2 , γ 3 and γ 12 are the surface energy of the solid/vapour, liquid/vapour and

solid/liquid interfaces, respectively. The contact angle (θ) is basically a parameter used to
explain the wetting interaction between a liquid and a solid.
In theory, the contact angle may vary from 0o to 180o. If the molecules of a liquid
have a stronger attraction to the molecules of a solid surface than to each other, wetting of the
surface occurs and vice versa. Accordingly, surfaces having θ > 90o are called hydrophobic
and hydrophilic when θ < 90o.
Contact angle measurements with different liquids of known surface tension values
constitute the basis for the determination of the surface energy of a solid. Nevertheless, the
resulting surface energies may differ depending on a specified approach which has been
followed. The most commonly used approaches are listed in Table 2.1. The main difference
between them is the number of different test liquids required for a single parameter
determination (Sharma and Roa, 2002). The main concepts of each approach are briefly
highlighted below.

• Zisman approach
Zisman and his co-workers (Fox and Zisman, 1950; Zisman, 1964) introduced an
empirical approach to estimate the surface energy of a solid surface from contact angle data.
This approach is limited to systems where mainly dispersive forces interact, i.e. it is not
applicable to polar systems (Karbowiak et al., 2006). In addition, the approach cannot be
adopted since different values of γ 2 can be obtained for a particular solid surface, depending
on the specific liquids that are used (Sharma and Rao, 2002).

• Fowkes approach
Fowkes (1964) proposed a model where the surface energy of a solid surface can be
considered as a sum of dispersive (γd) and non-dispersive components (γp). He attributed the
(γd) term only to the London dispersion interactions and (γp) due to hydrogen bonding and
metallic bonding. Most of these forces, such as metallic bonding and hydrogen bonding, are a
function of specific chemical nature. However, London dispersion forces exist in all types of
matter and always give an attractive force between adjacent atoms or molecules. Therefore,
this approach can preferably be used when the interacting forces are entirely dispersion forces

15
2 Literature Review

such as high energy liquids (e.g. water) in contact with low energy solids (e.g. Teflon-like
materials).

• Owens, Wendt or geometric mean approach


This approach was originally proposed by Girifalco and Good (1957) and later
modified by Owens and Wendt (1969). They divided the total surface energy into two
components of dispersive and hydrogen bonding components. From practical use, the
approach is not straightforward since different testing liquids yield different values for the
polarities of a given solid surface. This arises from the fact that the polar components of the
surface tension in this approach are, in most cases, asymmetric electron donor (γ−) and
electron acceptor (γ+) contributions which cannot be described by the usual geometric mean
method.

• Wu or harmonic mean approach


A further possibility of combining the polar and dispersive components had been
developed by Wu (1971 and 1973). In this approach, the surface energy between a liquid and
a solid can be evaluated by the harmonic mean approach (Eq. 2.4) by using at least two test
liquids with known surface tension data. Wu (1973) suggested that the harmonic mean
equation might be better for a polar system than other previous approaches. It is especially
applicable for low energy systems such as polymers.

• Equation of state approach


Neumann et al. (1974) derived an empirical equation for polymer systems. Here the
surface energy of a solid surface can be determined by measuring the contact angle of a
single liquid with a known surface energy. The disadvantage of this approach is that only
total surface energy can be estimated without any details on the dispersive and polar
components (Lee, 1993).

• Lewis acid-base approach


The acid-base approach was developed by van Oss et al. (1987 and 1988). They
expressed the total surface energy in terms of a non-polar component or Lifshitz-van der
Waals component (γLW) and a polar component or Lewis acid-base component (γAB). Through
contact angle measurements with at least three liquids with known surface tension

16
2 Literature Review

components γ 3 , γ 3 and γ 3 , the surface energy components of a solid surface can be


LW + −

determined.
The choice between these different approaches is still a matter of much debate and
studies performed in different fields have concluded that, the choice of a specific approach
mainly depends on the system under consideration (Ginl et al., 2001). Sharma and Rao
(2002) showed the usefulness of the Lewis acid-base approach, since it provides detailed
information about the modified surfaces. In addition, in this approach the electron donor
component allows the categorization of different modified surfaces with similar qualitative
composition. Accordingly, the Lewis acid-base approach is selected in this study for
determination of the surface energy of modified surfaces.
Having examined the concept of surface energy and its measuring approaches, it is
therefore essential to address the common techniques used to modify the surface energy.
Many different modification technologies are available that can be classified into 1) external
coatings to the substrate including PTFE, electroless Ni-P-PTFE, silica coatings or nano-
coatings and 2) embedded coatings such as ion implantation and sputtering. A detailed insight
into various modification techniques is given by Zettler (2002).

17
2 Literature Review

Table 2.1 Various approaches to determine surface energy

Name of approach Mathematical formula Comments


Zisman, (1950-1964) Graphical estimation - at least three test liquids should be used
- not applicable to polar systems

Fowkes, (1964)
γd - a single non-polar liquid with a known γ 3d surface energy is required
cos θ = 2 γ . 3 − 1
d
2 (2.2)
γ3 - applicable for high energy liquids in contact with low energy solids

Owens and Wendt, - two test liquids (one of them should be a polar) are required
γ 3 (1 + cos θ ) = 2 γ 2d γ 3d + 2 γ 2p γ 3p (2.3)
(1969)

Wu, (1971 and 1973) 4γ 2d γ 3d 4γ 2p γ 3p - applicable for low energy systems


γ 3 (1 + cos θ ) = +
18

(2.4)
γ 2d + γ 3d γ 2p + γ 3p - two test liquids (one of them should be polar) are required

Equation of state, o ( γ 3 −γ 2 )
2
- applicable only for dispersive interactions
γ 3 (1 + cos θ ) = 2 γ 2 ⋅ γ 3 ⋅ e −C (2.5)
(1974) - a single liquid with a known surface energy is required

Lewis acid-base, (1987 - at least two polar test liquids with one non-polar liquid should be used
γ 3 (1 + cosθ ) = 2( γ 2LW ⋅ γ 3LW + γ 2+ ⋅ γ 3− + γ 2− ⋅ γ 3+ ) (2.6)
and 1988)

18
2 Literature Review

2.1.2.2 Intermolecular interaction energies


The state in which two dissimilar bodies are held together by interfacial energies is
called adhesion. The interfacial energies holding them may arise from chemical bonding,
electrostatic interactions, mechanical interlocking. If two solid surfaces of 1 and 2 are
contacted in a third medium 3, as shown in Figure 2.4, the energy required to separate them
from equilibrium position to infinity is known as the adhesion work. It is directly
proportional to the total interaction energy.

W132

W132 = W12 + W33 − W13 − W23

Fig. 2.4 Work of adhesion in a third medium (Israelachvili, 1991)

Based on the DLVO theory (Derjaguin and Landau, 1941; Verwey and Overbeek,
1948), the intermolecular interaction energies between two bodies mainly consist of:

1. The Lifshitz-van der Waals energy ( ΔE132


LW
), which depends on the geometry and
physical/chemical properties of the interacting bodies. The resulting interaction laws for some
common geometries are summarized by Israelachvili (1991). The most common interacting
geometry for fouling can be very often approximated to a spherical-plate type, yielding:

− A132 R
ΔE132
LW
= (2.7)
6H

A132 = ( A11 − A33 )( A22 − A33 ) (2.8)

where R is the crystal radius, H is the distance of separation and A is the Hamaker constant
related to the properties of the interacting materials. van Oss (1994) presented a very simple

19
2 Literature Review

method for the calculation of the Hamaker constant based on the surface energy of the
interacting materials:

Aii = 24π H o2 γ iLW (2.9)

where Ho is the minimum equilibrium distance between the two interacting bodies. Thus,

ΔE132
LW
=−
4π H o2 R
H
(γ 1
LW
− γ 3LW )( γ LW
2 − γ 3LW ) (2.10)

Here, the van der Waals interaction energy is largely insensitive to variation of
electrolyte concentration and pH (Israelachvili, 1991).

2. Electrostatic double layer energy ( ΔE132


EL
). When a solid surface is immersed in a solution
containing dissolved ions, it shows a general tendency to acquire an electrical surface charge
either by ionization or dissociation of surface groups or by adsorption of ions from solution
(Oliveira, 1997). This creates an ionic cloud extending into the solution, which is called the
electric double layer (electrostatic screening). The electric double layer can be divided into
two layers:
i) an immobile layer which is called the Stern or Helmholtz layer. The ions of this
layer are strongly bound to the surface, and

ii) a diffuse layer which consists of free ions that move in the fluid under the influence
of an electric potential difference between the fluid bulk and the solid surface (Dukhin
and Derjaguin, 1974).

⎡ 2ζ ζ 1 + exp(− κ H ) ⎤
ΔE132
EL
= π ε R (ζ 12 + ζ 22 ) ⎢ 2 1 22 ln + ln{1 − exp(− 2κ H )}⎥ (2.11)
⎣ ζ 1 + ζ 2 1 − exp(− κ H ) ⎦

Interactions between these ionic clouds lead to the generation of long-range


electrostatic interactions that can be repulsive or attractive, depending on whether interactive
surfaces have like or opposite charge, respectively.

20
2 Literature Review

The van der Waals attraction energy always exceeds the double-layer repulsion
energy at small enough distance since it is a power-law interaction (i.e., ΔE132
LW
α − 1 / H n ),
whereas the double-layer interaction energy remains finite as H → 0 (see Figure 2.5)
(Israelachvili, 1991).
energy barrier

double layer weak adhesion


Total interaction energy

repulsion energy

0 5 10 Distance, H (nm)

van der Waals


attraction energy

strong adhesion

Fig. 2.5 Schematic energy versus distance profiles of DLVO interaction (Israelachvili, 1991)

Later on, van Oss (1994) extended the DLVO theory to include:

3. Lewis acid-base interaction energy ( ΔE132


AB
). When a solid that has electron donor and
electron acceptor capacities is immersed in water (which has strong electron donor and
electron acceptor capacities) a strong acid-base attraction will occur (van Oss et al., 1986).
Accordingly, the Lewis acid-base interaction energy has a polar character and can be either
attractive or repulsive (van Oss, 1994). It is expressed by:

⎛ Ho − H ⎞
ΔE132
AB
= 2π R λ ΔE132
AB
( H o ) exp⎜ ⎟ (2.12)
⎝ λ ⎠

⎛ γ +γ − + γ +γ − − γ +γ − + ⎞
⎜ 1 3 2 3 3 3 ⎟
⎜ − + ⎟
ΔE132
AB
(Ho ) = 2⎜ γ1 γ 3 + γ 2 γ 3 − γ 3 γ 3 −⎟
− + + −
(2.13)
⎜ + − ⎟
⎜ γ 1 γ 2 − γ 1− γ 2+ ⎟
⎝ ⎠

21
2 Literature Review

where λ is the correlation length pertaining to water molecules. It approximately equals 0.2
nm for pure water (Oliveira, 1997).

4. Brownian motion ( ΔE Br ) will bring particles into contact with the surface in question.

Adhesion occurs only if the flow of the liquid is not able to provide a shear force strong
enough to overcome the attractive forces between the particulate matter and the solid surface.
Particles adhering to a surface have two instead of three degrees of freedom as the one
perpendicular to the surface is blocked by bonding. Since Brownian motion comprises (1/2
kBTb) per degree of freedom, the corresponding free energy term of a particle adhering to a
surface equals 1kBTb.
Thus, the total interaction energy ΔE132
TOT
between a deposit and a solid surface can be written
as the sum of the respective interaction energies (van Oss, 1994; Oliveira, 1997):

ΔE132
TOT
= ΔE132
LW
+ ΔE132
EL
+ ΔE132
AB
+ ΔE Br (2.14)

Intermolecular interaction energies

DLVO theory

Lifshitz-van der Waals Electrostatic double layer Lewis acid-base Brownian


interaction energy, interaction energy, interaction energy, motion,
ΔE132
LW
ΔE132
EL
ΔE132
AB
ΔE Br

Extended DLVO theory

Fig. 2.6 Molecular interactions energies influencing adhesion

The balance between all interaction energies in terms of ΔE132


TOT
determines whether

the deposits will attach on the surface or not. For instance, adhesion takes place when ΔE132
TOT

is negative implying that fouling reduction can only occur when ΔE132
TOT
increases. It is
therefore indispensible firstly to determine all intermolecular interaction energies formulated

22
2 Literature Review

in Eq. (2.14) then to identify what intermolecular interaction energy and also which surface
energy component are dominant for the calculation of ΔE132
TOT
. Thus as a preliminary step, it is
essential to review the previous studies that have been carried out to investigate the influence
of surface energy components on fouling propensity.

2.1.2.3 Effect of different surface properties on fouling mechanisms


There is ample evidence to demonstrate that surface energy can significantly
influence the fouling process. In most cases, but certainly not all, lower surface energy
materials reduce fouling adhesion compared with untreated surfaces (Müller-Steinhagen and
Zhao, 1997; Zhao et al., 2005b). Table 2.2 presents a summary of previous studies on the
effect of surface properties on fouling mechanisms. The Table is not limited to studies on
crystallization fouling but includes also studies on biofouling and dairy applications. As the
focus in this study is on crystallization fouling, only the respective literature will be discussed
here.
In the past several years, many attempts have been made to assess the effect of surface
energy properties on calcium sulphate fouling mechanisms under various heat transfer modes.
A pioneering study by Müller-Steinhagen and co-workers showed that the deployment of
SiF3+ implanted surfaces during pool boiling of CaSO4 solutions causes the final heat transfer
coefficient to remain 1.5-3 times higher than that of the original stainless steel surface. In
addition, the formed deposit was loose and porous in nature and broke off more easily than
the dense deposit formed on the original surface (Müller-Steinhagen and Zhao, 1997). Other
implantation surface such as H, F and Si as well as diamond like carbon (DLC), ta-C and ta-
C:F coated surfaces improved anti-fouling effects, since i) the final heat transfer coefficients
were about 2-3 times higher than those obtained for the original surface, and ii) the deposit
formed was easy to remove suggesting lower deposit adhesive strength (Bornhorst et al.,
1999).
The influence of surface energy during forced-convective heat transfer is somewhat
different to that for pool boiling. In most cases, the surface energy reduction corresponded to
a considerable reduction of the fouling resistance as a result of different deposit structure on
various modified surfaces (Förster and Bohnet, 1999; Zettler, 2002). However, despite the
fact that the low surface energy of ion implantation, ion sputtering, carbo-nitriding and
oxidizing and Ni-P-PTFE modified surfaces leads to less deposit formation, no simple

23
2 Literature Review

correlation between surface energy and CaSO4 fouling behaviour can be found (Zettler et al.,
2005). On the other hand, the lowest energy surfaces, e.g. SiF3+ implanted and DLC-sputtered
plates do not exhibit the highest reduction in fouling resistance (Zettler et al., 2005). Later,
Augustin et al. (2005) tailored DLC-coatings by variable addition of elements such as Si and
O, and investigated them under convective heat transfer to CaSO4 solutions. They found that
the induction time was prolonged approximately four times compared to the stainless steel
surface. However, this propensity was not directly related to the surface energy as again no
clear-cut relationship between surface energy and fouling behaviour can be established
(Augustin et al., 2005). A similar assertion was drawn by Geddert et al. (2009). They showed
that the comparison of total surface energy with induction time yields no correlation between
surface energy and fouling behaviour.
In addition to the previous discussion, experimental results reported to-date do not
provide a consistent picture. For instance, it has been found that the polar component of the
surface energy plays an important role in deposit adhesion on a metal substrate (Müller-
Steinhagen et al., 1997). A similar assertion was drawn by Augustin et al. (2007). They found
that the induction time increases with increasing polar component of surface energy. In
contrast, Geddert et al. (2009) found that the comparison of polar and disperse components
with fouling did not yield any
correlation. Another interesting result is 4

that the slope of the nucleation curve


slope of nucleation curve

3
⎛ d (ln(t ind )) ⎞
⎜⎜ −2
⎟⎟ of CaSO4 has a linear
fluorine

⎝ d (ln ( S )) ⎠
hydrogen

2
relationship with the electronegativity of
oxygen

the modified surfaces (H, F, O and neon 1

ion implanted surfaces) as shown in


neon

Figure 2.7 (Rizzo, 2008). The literature 0


-100 -50 0 50 100 150 200 250 300 350 400
review, therefore, underlines that no electronegativity (Eea) [kJ/mol]

satisfactory and simple correlation is yet


available between the impact of surface
energy and fouling propensity.

24
2 Literature Review

Table 2.2 Summary of previous studies on various fouling mechanisms of different modified surfaces

Reference Solution Heat Surface Operating conditions Conclusions


transfer cb, v, m/s Tb, Ts,oC
o
mode g/L C q&
Müller- CaSO4 pool boiling +
SiF ion implanted
3
1.2- ----- ---- - final heat transfer coefficient remains 1.5-3 times above that
Steinhagen and surfaces 1.6 - of the original stainless steel surface.
100-200
Zhao, (1997) - loose and porous deposit formed in comparison to packed
kW/m2
deposit on the original surface.
Förster and CaSO4 convective DLC, PTFE, Al, 2.5 0.2 42 75 - polymer surfaces are not well suited for heat exchangers
Bohnet, (1999) heat transfer Cu and SS 31.8 kW/m2 exposed to liquid salt solutions of inverse solubility.
Förster et al. CaSO4 convective SS, DLC, 2.5 0.2 42 75 - negate a direct relation between surface energy and fouling
(1999) heat transfer PTFE, FEP, PFA, tendency with respect to induction time.
31.8 kW/m2
Cu and Al
Bornhorst et CaSO4 pool boiling F, H, S implanted 1.2- ----- 100 - the final heat transfer coefficients are about 2-3 times higher
al. (1999) surfaces; ta-C and 1.6 than for the untreated stainless steel surface.
100-400
ta-C:F - deposit formed was easy to remove.
kW/m2
Zhao et al. CaSO4 flow boiling Ni-P-PTFE 2 0.8 200 kW/m2 - low energy surfaces have the potential to reduce mineral
25

(2002) fouling.
Augustin et al. CaSO4 convective SS, DLC, SICAN, 3.4 ----- 42 75 - no clear-cut tendencies between surface energy and
(2005) heat transfer SICON, CrN, sol- induction time.
gel - the induction time could be extended four times by tailoring
the DLC coating with silicon.
Zettler et al. CaSO4 convective F, H, Si ion 3 0.35 51. 86 -no correlation between surface energy and fouling behaviour.
(2005) heat transfer implanted, DLC 5
and PTFE
Zhao et al. CaSO4 pool boiling Ni-P-PTFE 1.2 ----- ---- - the surface energy of coatings had a significant influence on
(2005b) - 100 kW/m2 the adhesion of CaSO4 deposits.
- when the surface energy of the coatings was around 28
mJ/m2, the adhesion of CaSO4 deposits was minimal.
Rizzo, (2008) CaSO4 convective F, H, O and neon 1.6-2 0.125 50 97 - fouling tendency could not be correlated with surface
heat transfer implanted surfaces -2 energy.
-the induction time increases linearly with electronegativity.

25
2 Literature Review

Geddert et al. CaSO4 convective SS, DLC, SICAN, 3.4 ----- 42 75 - no correlation between surface energy and fouling
(2009) heat transfer SICON and CrN behaviour.
- the combination of electrochemical treatment and PECVD
coating is able to increase the induction time.
Liu and Zhao, bacteria convective Ni-P-PTFE ---- ----- ---- ----- - the total interaction energy increases with decreasing surface
(2005) heat transfer energy.
Zhao et al. bacteria convective Ni-Cu-P-PTFE 1.2 - when the surface energy was in the range 26–30 mJ/m2, the
(2005a) and heat transfer adhesion of both bacteria and CaSO4 deposit was minimal.
CaSO4 - increasing PTFE content causes surface energy to decrease.
100 kW/m2
Visser (2001) whey convective DLC, PTFE, ion ---- ----- ---- ----- - the Lewis acid-base energy has a high contribution to the
protein heat transfer implanted total interaction energy.
and - an important parameter to reduce fouling is to increase the
calcium γ 2− of the steel surface.
phosphate
Santos et al. whey convective SS, SiF3+ and 3 ----- 85 ----- - no simple relation between surface energy and protein
(2003) protein heat transfer adsorption is apparent.
MoS22+ ion - the DLC sputtered surfaces were most promising in terms of
26

implanted, DLC reducing protein adsorption.


sputtered surfaces
Augustin et al. whey convective DLC (a-C:H), 3- ----- 50 70-80 - the induction time increases with increasing polar
(2007) protein heat transfer (a-C:H:F), 3.5% component of surface energy.
(b-C:H:Si:O), Cu,
and Al
Wu and calcium convective TiO2 ---- ---- ---- ----- - the Lewis acid-base energy plays a crucial role in the
Nancollas, phosphate heat transfer nucleation of calcium phosphate
(1998)
Rosmaninho et calcium convective SiF3+ , MoS22+ , ------ ------- 40- -------- - surfaces having higher γ 2− are more prone to calcium
al. (2005) phosphate heat transfer 75
SiOx, DLC-Si-O, phosphate deposition.
Ni-P-PTFE, and
Silica coated
surfaces
Rosmaninho et calcium convective TiN sputtered 44 70 - high γ 2− surfaces are less prone to ion-controlled reaction.
al. (2007) phosphate heat transfer surfaces
- high γ 2− surfaces are more prone to particle adhesion.

26
2 Literature Review

2.1.2.4 Fouling models with consideration of surface energy properties


The purpose of any fouling model is to assist heat exchanger engineers to make a
reliable prediction of fouling propensity in order to optimally design and operate heat
exchangers. Kern and Seaton (1959) proposed a pioneering fouling model as:

dM d ⎛ dm ⎞ ⎛ dm ⎞
=⎜ ⎟ −⎜ ⎟ (2.15)
dt ⎝ dt ⎠ d ⎝ dt ⎠ r

⎛ dm ⎞ ⎛ dm ⎞
where ⎜ ⎟ and ⎜ ⎟ are the deposition and removal rates, respectively. Subsequently,
⎝ dt ⎠ d ⎝ dt ⎠ r
the model served as a basis for the development of many new correlations usually based on
several assumptions, e.g. only a single fouling mechanism and a homogenous fouling layer
are considered. Furthermore, the models only included a few parameters such as velocity,
concentration and temperature from the many parameters that affect practical fouling
processes such as those for crude oil. Most notably there are only a few models that include
surface related properties.
One of the first semi-empirical models that incorporated the effect of surface energy
on dairy fouling processes was presented by McGuire and Swartzel (1989). It is based on the
influence of the solid surface energies on protein adsorption from milk. It has been observed
that the net amount of protein deposition follows an Arrhenius type expression of the general
form:

∂M ⎛ Ea ⎞
= K exp⎜ − ⎟ (2.16)
∂t ⎜ R g Ts γ 2 − γ 2,min ⎟
⎝ ⎠

E a = 223.515Ts + 2805.711γ 2 − γ 2,min − 78.087Ts γ 2 − γ 2,min

γ 2,min = 30.72 + 0.003Ts


100oC < Ts < 158oC

where M represents the mass of protein and carbohydrate adsorbed at any time onto a
substrate with a surface temperature (Ts) and a surface energy ( γ 2 ); γ 2,min represents a

material evoking minimal biological adhesion and has been expressed as a function of surface

27
2 Literature Review

temperature. It has been found that the activation energy (Ea) is increased as γ 2 approaches
γ 2,min , and consequently the amount of proteinaceous deposition is reduced. In addition, a

significant reduction of activation energy has been observed at constant γ 2 with increasing
surface temperature, which in turn leads to increase the deposition rate. Later, Förster and
Bohnet (2000) developed an interfacial defect model based on experimental results that
would predict the optimal choice of surface martial. The model is capable of relating wetting
characteristics to the adhesive strength ( ξ ) as defined by Wu (1982).

−1
⎛ λ ⎞
ξ = E ⎜⎜1 − 12 ⎟⎟ (2.17)
⎝ γ2 ⎠

λ12 = γ 2 − γ 3 − γ 12 (2.18)

Here, an increasing value of spreading coefficient ( λ12 ) means an improved


wettability and consequently a higher adhesive strength. By combining Eq. (2.1) and Eq.
(2.18) one obtains:

λ12 = γ 3 (cos θ − 1) (2.19)

According to Eq. (2.19) two conclusions can be drawn:

i) If no spreading occurs corresponding to θ = 180o, Eq. (2.19) yields λ12 = −2γ 3 ,

while λ12 = 0 in case of spontaneous spreading, i.e. θ = 0o.


ii) In case of λ12 > 0, the system will no longer be in a state of equilibrium.

Förster and Bohnet (2000) found that the interfacial defect model is valid for metallic
heat transfer surfaces, whereas the deployment of polymeric surfaces leads to a significant
deviation of the corresponding data points from the regression curve (see Figure 2.8). Förster
and Bohnet (2000) attributed the performance of polymeric surfaces to the influence of
surface topography on precipitation fouling. They stated that, due to the coating procedure, a
polymeric surface tends to have more critical flaws which serve as regions for preferred
nucleation and consequently reduced induction time.

28
2 Literature Review

350
Lewis acid-base approach
Ms(II)
Metallic heat transfer surfaces
300
Polymeric heat transfer surfaces

250
induction time [h] PFA
200
FEP
Ms(I)
150 SS

100 PTFE Bronze

50 Al
Cu

0
-50 -40 -30 -20 -10 0 10
2
spreading coefficient (λ 12) [ mJ/m ]

Fig. 2.8 Induction time versus spreading coefficient (Förster and Bohnet, 2000)

Subsequently, Förster (2001) found that the fouling propensity during the induction
1
period is affected by two parameters, i) surface texture parameter ( K f = ) and ii)
RZ Rq

⎛ λ ⎞
surface energy parameter ⎜⎜ Λ = 1 − 12 ⎟⎟ as shown in Eq. (2.20). To mitigate fouling,
⎝ γ2 ⎠
maximum values of Kf and surface energy parameter have to be attained.

[
t ind = 6.4 exp (27.7 Λ − 34.5)
0.243
] ⎡⎢1 + 5.38⎛⎜⎜ 10 Km
10
f
−2
⎞⎤
⎟⎟⎥ (2.20)
⎢⎣ ⎝ ⎠⎥⎦

To evaluate the contribution of interfacial energy to membrane fouling, Kwang-Ho


and Chung-Hak (2000) suggested that the interfacial energy ( ΔG f ) is governed by three

factors; γ 3d , γ 1d and γ 2d as described in Eq. (2.21). Here, γ 1d depends on the solutes in the

aqueous feed solution and γ 2d on the membrane materials to be used.

ΔG f = a + b γ 2d (2.21)

29
2 Literature Review

a = 2 As (γ d
1 γ 3d − γ 3 ) (2.22)

b = 2 As ( γ 3d − γ 1d ) (2.23)

Kwang-Ho and Chung-Hak (2000) found that the dispersion force ( γ 1d ) of the solute
(anaerobic biological mixed liquor) is stronger than that of water molecules. This is due to the
high polarity of water molecules (Wu, 1999). Accordingly, the “b” term in Eq. (2.23) would
be negative and consequently ΔG f decreases with the increase in γ 2d . Under such

circumstances, membrane fouling is expected to be more severe if the γ 2d value increases


with a change of membrane material. The prediction of fouling tendency according to Eq.
(2.21) was in better agreement with the experimental results than that based on surface
hydrophobicity (Kwang-Ho and Chung-Hak, 2000)
Recently, it has been found that under certain conditions some systems may foul, even
though the total interaction energy ΔE132
TOT
between deposit and metal surface is positive. For

instance, if the substantial cohesive energy ΔE131


TOT
between the foulant particles leads to

coagulation into larger particles, this may affect the adhesion energy ΔE132
TOT
to become
negative as the attractive energy. This is because both Lifshitz-van der Waals and Lewis acid-
base interaction energies increase significantly with increasing particle diameter according to
Eqs. (2.10) and (2.12). In addition, if the cohesive energy ΔE131
TOT
is equal to that of the

adhesion energy ΔE132


TOT
, then the energy characteristics of the heat transfer surface will be the
same as that of the particles. When this happens then the colloidal particles in the near-wall
boundary layer may not attach to the surface but remain suspended in the solution in some
sort of dynamic equilibrium. Kwok and Neumann (1996) found that the Lewis acid-base
approach does not consider the cohesive interaction between foulant particles. Accordingly,
Zhao and Müller-Steinhagen (2002) assumed that the calculation of both cohesion and
adhesion energies should be only dependent on the Lifshitz-van der Waals energy. Based on
this hypothesis they developed the following criterion to determine whether a system will
foul or not:

ΔE131
TOT
− ΔE132
TOT
=
24 π H o2 R
12 H
(γ 1
LW
− γ 3LW )( γ 1
LW
+ γ 3LW − 2 γ 2LW ) (2.24)

30
2 Literature Review

They proposed that if

• ΔE131
TOT
− ΔE132
TOT
> 0, then fouling occurs

• ΔE131
TOT
− ΔE132
TOT
= 0, then foul will not occur

If the system does not foul, i.e. ΔE131


TOT
− ΔE132
TOT
= 0 a value of surface energy γ 2,min
should exist at which fouling is minimal:

⎛1⎞
(
γ 2,min = ⎜ ⎟ γ 1LW + γ 3LW
⎝2⎠
) (2.25)

In addition to this assumption, Zhao and Müller-Steinhagen (2002) hypothesized that


the induction time and asymptotic fouling resistance are related to the minimum surface
energy as per Eqs. (2.26) and (2.27)

100
tind = (2.26)
γ LW
2 − γ 2,min

R*f = B1 ⎛⎜ γ 1LW − γ 3LW ⎞⎟ exp⎛⎜ − B2 / γ 2LW − γ 2, min ⎞⎟ (2.27)


⎝ ⎠ ⎝ ⎠

The literature survey presented above highlights the complexity of incorporation of


surface related properties into any fouling model, i.e. there is still no generally valid
correlation to predict the effects of surface energy on the fouling behaviour. Much of the
difficulty stems from the inability to distinguish the dominant specific parameters and the
lack of understanding of interaction between the deposits and the heat transfer surface. It can
also be concluded that the topic of altering the surface properties to mitigate fouling remains
almost unexplored. Furthermore, there is no consensus among investigators on the effect of
surface energy as well as of the interaction energies on fouling propensity. This has led
various investigators to come up with contradictory results for different foulants such as
bacteria, calcium phosphate and calcium sulphate. Part of these contradictory results must be
attributed to the unsystematic work that has been carried out to investigate the individual
interfacial energies in the DLVO theory. Accordingly, there are still several open questions
that should be addressed through more rigorous experimental and theoretical studies such as:

31
2 Literature Review

i) there is little available information to assess the contribution of the interfacial


energies to the fouling process or on the relative predominance of one of these
energies over the other with respect to the formation of the initial fouling layer
and,
ii) there is still no concrete model to relate surface interfacial energies to fouling. If
available, such models are restricted to a small data base and to the assumptions
that have been made to derive these models.

32
3 Experimental Set-Up and Procedure

Chapter 3
EXPERIMENTAL SET-UP AND PROCEDURE

The fouling experiments of this study are aimed at simulating the conditions that were
likely to occur in a plate heat exchanger that has to be utilized within the framework of the
MEDESOL project. The experimental work proceeded as follows: i) laboratory tests of
various structured and coated surfaces; ii) based on these preliminary findings, the coating
company is asked to coat the plates of compact heat exchanger; iii) field examination of
coated plates in a pilot desalination plant in Almeria, Spain and removal of plates after
several fouling runs, and v) repetition of all preceding steps if required. In parallel, laboratory
experiments are a pivotal part of the MEDESOL project as they would facilitate better
understanding of the mechanisms which influence heat transfer fouling under convective heat
transfer conditions.

3.1 Experimental Apparatus


3.1.1 Test rig
A picture of the experimental set-up is shown in Figure 3.1 and a schematic diagram
in Figure 3.2. The complete rig is made from stainless steel, except for the connections to the
pump, which are of ethylene propylene diene M-class rubber (EPDM), a chemical resistant
rubber which is used to isolate the vibration of the pump from the rest of the set-up. The
closed loop set-up (see Figure 3.2) consists of two identical test sections operated in parallel.
Each test section is a vertical rectangular duct. On both wide sides of each duct, openings
with a dimension of 59 mm x 50 mm exist where the specimens [5] can be mounted. The
heated zone of the specimen is positioned 1.5 m upstream of the duct inlet. This length allows
flow to reach fully developed conditions. Using a centrifugal pump [2], the solution is
pumped from a 47 liter stainless steel tank [1] through a polypropylene-polyethylene filter [3]
to remove any broken crystals and impurities from the solution, which might artificially act
as additional nucleation sites. Subsequently, the flow is divided into two equal streams to the
left and right ducts as shown in Figure 3.2. The flow rate is kept constant and controlled
separately using two magnetically inductive flow meters [4] in conjunction with PID
automatic controllers and two electrically controlled valves [6]. The two partial flows are
recombined after the heated zone and then fed back to the supply tank. To avoid any damage

33
3 Experimental Set-Up and Procedure

to the pump and to have better flow control, a bypass pipe system from the pump back to the
tank has been installed. The temperature of the solution in the tank is regulated using a water-
cooled coil inside the tank, together with two heater pads [7] installed on the outer surface of
the tank.

rectangular duct
test section

specimen holder specimen-front view

Fig. 3.1 Experimental set-up

34
3 Experimental Set-Up and Procedure

6 6

Sample holder

Sample holder

Sample holder

Sample holder
M M
TR TR TR TR

o o o o
o o
5 o o 5

TR TR TR TR

Test section R
Test section L
FIC FIC
cooling

M
water

4 4
1
M TR
heater

7 47
60 L L

3
Main tank
Filter,12 µm

5 m 3/h
2

Fig. 3.2 Schematic diagram of the experimental set-up

3.1.2 Heat transfer specimen


The heat transfer specimen employed in this study is of 0.3 mm high-grade stainless
steel sheet metal AISI 304 BA with 50 mm × 59 mm area, on which two copper blocks (40
mm × 20 mm × (10 and 15) mm) are vacuum-soldered as depicted in Figure 3.3. Six K-type
(Ni-Cr / Ni-Al) thermocouples are inserted into the copper blocks to measure the temperature
gradient, from which the surface heat flux can be calculated. Two 450 Watt golden infrared
heaters (Heraeus Noblelight GmbH, Hanau, Germany) are used to provide the required heat
flux. The maximum attainable heat flux at the specimen is 150 kW/m2. During each
experiment, the transmitted heat flux of all specimens is sequentially measured and regulated
by means of PID automatic controller dimmers with the maximum deviation from the desired
value being less than ± 4.8%. Each individual specimen is fixed in a special holder by means
of a vacuum pump. Afterwards, the holder is fastened to the test section with six bolts. A
cross-section of the sample holder including the IR radiators and a view of the flow channel
is depicted in Figure 3.4

35
3 Experimental Set-Up and Procedure

copper blocks thermocouples thermocouples


50 mm

sample holder
sample holder
40 mm
q q
substrate
59 mm

20 mm

vacuum line
vacuum line
thermocouple
position 2
thermocouple
position 1 SS

Fig. 3.3 Heat transfer specimen Fig. 3.4 Holder of the heat transfer specimen

The surface temperature of the specimen is determined using the following equation:

q& x
Ts = Tth − (3.1)
k

In this equation, the thermal resistance “x/k” between the measuring point of the
thermocouples (position 1 in Figure 3.3) and the heating surface should be obtained.
Although the distance between the thermocouple and the heat transfer surface is known, the
determination of x/k is not quite simple due to the changing thermal conductivity arising from
the presence of the soldering material. Hence, the Wilson Plot procedure is followed for
experimental determination of the x/k values (see Appendix A1 for more details).

3.1.3 Data acquisition and reduction


A fully-controlled data acquisition system is used to log and process all experimental
data such as solution flow rate, bulk temperature, heat flux, cooling water temperature and
surface temperature. A graphical programming language “VEE pro 6” (Agilent Technologies
GmbH, Germany) is used for data acquisition in specified time intervals. The analysis of the
measured data takes place on a separate PC where the outcomes are stored in an Excel sheet.

36
3 Experimental Set-Up and Procedure

For better evaluation of the system, the measured temperature time series are plotted as a
graphical view with a time interval of 26 seconds.

3.2 Surface Characterizations


Fundamental understanding of important surface properties such as roughness, contact
angle and surface energy which would influence fouling mechanisms requires accurate
quantification of these parameters. It is, therefore, necessary to elucidate the techniques and
devices that are employed for the measurement of these properties.

3.2.1 Surface roughness


The first step in surface characterization in the present study is to measure the
roughness index of the specimens. A stylus instrument (Perthometer M4Pi, Mahr GmbH,
Göttingen, Germany) as illustrated in Figure 3.5 is used to measure surface roughness.
Surface roughness can be expressed in different terms, e.g. the roughness profile (Ra), the
mean roughness depth (Rz) and the root mean square (Rq).

Filters Data processing


Tracing arm

Measuring amplifier

Sample

Fig. 3.5 Experimental arrangement of the stylus device

The roughness profile is defined as the arithmetic average deviation of the surface pits
and peaks from the mean line as described in Eq. (3.2) and Figure 3.6a. In practice, however,
roughness is not an easy quantity to measure. It can change with respect to space or direction,
and is not a scale-invariant. For the accurate measurement of roughness profile, a straight
pick-up with a constant speed is drawn over a length of 15 mm in three horizontal and
vertical lines. The pick-up has generated two dimensional images of the profile by assessing
the surface via the mechanical movements of a stylus tip. The overall mean roughness of the

37
3 Experimental Set-Up and Procedure

heat transfer specimen was determined by averaging the mean values of the 9 points at the
intersection between the horizontal and vertical lines, as shown in Figure 3.6b.

1 L
Y ( x ) dx
L ∫0
Ra = (3.2)

1 2 3 H1
profile high (y) [μm]

Ra 4 5 6 H2

7 8 9 H3

V1 V2 V3
trace length (x) [mm]

Fig. 3.6 a) Surface roughness profile as a function of distance and b) Positions and directions
of surface roughness

3.2.2 Contact angle


Two different approaches are commonly used to measure the contact angle of a solid
surface, 1) goniometric technique which involves the observation of a sessile drop of a test
liquid on a solid substrate and 2) tensiometric technique which involves measuring the forces
of interaction as a solid is contacted with a test liquid as in the case of using the Wilhelmy
plate method. In this study, the goniometric technique is used to determine the contact angles
by employing the dataphysics G10 contact angle measuring instrument as shown in Figure
3.7 in connection with a Drop Shape Analysis (DSA) program by Krüss GmbH, Hamburg,
Germany. The instrument consists of a charge-coupled device (CCD) video camera with a
resolution of 768 × 512 pixels, and a multiple dosing/micro-syringe with a needle of 0.5 mm
inner diameter.
Since surface contamination would influence the wetting behaviour of the solid
substrate (Mantel and Wightman, 1994), the specimens are ultrasonically cleaned by acetone

38
3 Experimental Set-Up and Procedure

(96%) and deionized water in sequence. The specimens are then dried in an air stream to
ensure that no acetone remained on the surface. Afterwards, a droplet of a test liquid is placed
on the surface under controlled room temperature. The drop image is then recorded by a
video camera and displayed on a monitor. Once stabilized, a frozen picture of the droplet will
be taken and the contact angle measured using the DSA program with an accuracy of ±0.1o.

syringe

droplet
CCD - camera
Heat surface
Light source

Fig. 3.7 Schematic diagram of the sessile drop for measuring contact angle

3.2.3 Surface energy


As explained in Chapter 2 in this study, the Lewis acid-base approach in accordance
with contact angle measurements is used to determine the surface energy components of the
modified surfaces. In this approach, the surface energy of a solid surface is composed of two
different types of interactions (Eq. 3.3); i) non-polar component ( γ iLW ) which refers to the

Lifshitz-van der Waals component, and ii) polar component ( γ iAB ) which consists of the
electron acceptor and electron donor components (Lewis acid-bases components) as shown in
Eq. 3.4 (van Oss et al., 1988).

γ i = γ iLW + γ iAB (3.3)

γ iAB = 2 γ i+ γ i− (3.4)

The solid/liquid interface energy is then given by

39
3 Experimental Set-Up and Procedure

γ 12 = γ 2 + γ 3 − 2( γ 2LW ⋅ γ 3LW + γ 2+ ⋅ γ 3− + γ 2− ⋅ γ 3+ ) (3.5)

The combination of Eq. (3.5) and Young Eq. (2.1) (Young, 1805) yields a relation
between the measured contact angle and the solid and liquid surface free energies as:

γ 3 ⋅ (1 + cos θ ) = 2( γ 2LW ⋅ γ 3LW + γ 2+ ⋅ γ 3− + γ 2− ⋅ γ 3+ ) (2.6)

The experimental procedure for the determination of surface components γ 2LW , γ 2+


and γ 2 has been described by Good et al. (1991). There are two methods to utilize Eq. (2.6):

i) to measure the contact angle of three polar liquids with known values of γ 3LW , γ 3+ and γ 3− ,

or ii) using only one non-polar liquid for finding γ 2LW and two other polar liquids to obtain γ 2+

and γ 2 . In this study, diiodomethane is used as a non-polar liquid, while water and ethylene

glycol as polar liquids. Table 3.1 lists the surface tension components of these liquids (Good,
1992; van Oss et al., 1988).

Table 3.1 Test liquids and their surface tension components


Liquids Surface tension components, mNm-1
γLW γAB γ+ γ- γ
Water, H2O 21.8 51.0 25.5 25.5 72.8
Ethylene glycol, C2H6O2 29.0 19.0 1.9 47 48.0
Diiodomethane, CH2I2 50.8 50.8 0 0 50.8

3.3 Experimental Procedure


3.3.1 Test solution and preparation procedure
As mentioned earlier in Chapter 1, calcium sulphate was selected as working fluid for
the fouling experiments in this study. In nature, calcium sulphate exists in three forms which
are different chemically by the amount of water contained in their structure (Palache et al.,
1951). The two common forms are calcium sulphate dihydrate (CaSO4.2H2O) and calcium
sulphate anhydrite (CaSO4). A third form is calcium sulphate hemihydrate (CaSO4.1/2H2O).
It is a metastable form and only precipitated when certain natural water is evaporated at a
temperature of 100oC. It has been found that the solubility of all three forms in pure water
decreases with increasing temperature (Patridge and White, 1929). However, pure calcium

40
3 Experimental Set-Up and Procedure

sulphate has a very low solubility in distilled water. In such a case, the calcium sulphate
solution for the experiments is prepared by mixing an aqueous calcium nitrate tetrahydrate
(Ca(NO3)2·4H2O) solution with an aqueous sodium sulphate (Na2SO4) solution in such a way
that the desired CaSO4 concentration is finally obtained.

Ca ( NO3 )2 .4 H 2 O + Na 2 SO4 → CaSO4 ( aq ) + 2 NaNO3 ( aq ) + 4 H 2 O (3.6)

These two chemicals are chosen because their higher solubility in water can provide a
high enough concentration of foulant ions in the solution (Najibi et al., 1996). The presence
of other ions, i.e. NO3− and Na 2+ in the solution increases the solubility of calcium sulphate.
This can be attributed to the electrostatic attraction between the sodium nitrate ions and the
ions with opposite charge formed by calcium sulphate (Marshall et al., 1964). For a given
concentration of calcium sulphate, the required mass of each chemical is weighted using a
digital scale type “Sartorius 1264 MP” and then dissolved separately in 23.5 liters of
deionized water and heated up to the final bulk temperature of 40oC. Afterwards, they are
mixed thoroughly for approximately three hours under constant temperature to ensure
homogeneity. A visual examination was done to ensure whether turbidity occurred or not as it
is a clear indication of precipitation. This arises particularly from a very high supersaturation
at elevated bulk temperatures. If turbidity occurred then the experiment should be terminated
and a new solution prepared. Once the mixing of salt solutions is accomplished, the calcium
concentration is determined by measuring the calcium ion concentration by potentiometric
EDTA (Ethylene-Diamine-Tetra-acetic Acid) titration (Fritz and Schenk, 1987). The
stoichiometric reaction is as follows:

(C a )
+ Indic − + H 2Y 2− → C a Y 2− + H Indic 2− + H + (3.7)

The titrant (H2Y2-) is a solution of the disodium salt of EDTA (disodium dihydrogen
ethylene-diamine-tetra-acetic acid). Its concentration is 0.01 mol/L. Eriochrome Black T is
used as the indicator and a certain amount of magnesium sulphate (MgSO4) with the same
concentration as titrant (H2Y2-) is also added, so that the added amount of magnesium is
subtracted from the total EDTA consumption. The pH value of the titration sample has to be
adjusted to a value of 9-10 by adding 2-3 ml of ammonia buffer solution with pH=10.

41
3 Experimental Set-Up and Procedure

3.3.2 Clean and fouling heat transfer coefficients


Each individual specimen is initially cleaned by toluene to remove any contamination
and then immediately mounted into the test rig. Afterwards the selected heat flux, bulk
temperature and fluid velocity are adjusted, and then the rig started to operate under clean
conditions with deionized water until steady-state is reached. This is reached when the
surface temperature remains constant. During this period, the test solution is prepared and
heated to the desired bulk temperature. If the surface temperature of all specimens and the
other test conditions are stable for more than two hours, the test run is stopped, the deionized
water removed and replaced by the test solution. The solution is then circulated for about 10
minutes with the heaters still off, to ensure that any dissolved air can find its way to
deaerators and leave the system. Afterward, the power supply to the specimen is switched on
and the data collection started.
During the experiment the Ca2+ concentration has to be determined by EDTA-titration
every 2-3 hours and adjusted by manually adding equimolar calcium nitrate tetrahydrate and
sodium sulphate to restore the CaSO4 concentration to the initial value. However, such
adjustment is less than 5% in the first 3 hours of the fouling runs. This is mainly because the
supply tank is relatively large which prevents rapid reduction of concentration when the
solution is continuously circulated. Furthermore, the chemical adjustment may lead to
increase the presence of other ions such as NO3− and Na 2+ in the solution which would cause
the solubility of calcium sulphate to increase (Marshall et al., 1964). Rigorous previous tests
in the same test rig found the effect of such ions on fouling behaviour is negligible (Rizzo,
2008). The adjustment process is carried out at identical times in each experiment in order to
obtain comparable results. Depending on the subsequent decreases in CaSO4 concentration
and on the deposition formation, this is not always possible. As soon as the surface
temperature of a specimen exceeded a set value, the respective heating has to be switched off.
After the runs, all heating lamps are switched off and the control software stops the test rig
automatically and stores all measured data in an output file.

3.3.3 Determination of basic fouling parameters


To monitor the impact of the modified surfaces on crystallization fouling, the
following parameters are used to evaluate the experimental results and to be the basis for
comparison between the modified surfaces.

42
3 Experimental Set-Up and Procedure

• Fouling resistance
The presence of the fouling layer on a heat transfer surface due to crystalline
deposition behaves as a barrier to heat transmission. This is due to the low thermal
conductivity of the deposit layer which severely reduces the effectiveness of the heat
exchanger. Figure 3.8 shows a typical clean and fouled heat transfer surface.

clean heat Tso fouling layer Tst


transfer surface Tf

Tb Tb

q&
q&

a) b)

Fig. 3.8 a) Clean heat transfer surface and b) Fouled heat transfer surface

By assuming one directional and constant heat flow, the total heat transfer rate from a
hot surface to a cold fluid for a clean surface involves only convection as per Eq. (3.8). For a
fouled surface the conduction resistance due to the presence of the fouling layer should be
considered as expressed by Eq. (3.9):

(Ts o − Tb )
q& = (3.8)
1
α

(Ts t − Tb )
q& = (3.9)
1
Rf +
α

From Eqs. (3.8) and (3.9) the fouling resistance at any time yields as:

43
3 Experimental Set-Up and Procedure

⎛ T − Tso ⎞
R f = ⎜⎜ s t ⎟⎟ (3.10)
⎝ q& ⎠

• Initial fouling rate


In this study, the initial fouling rate is considered as the slope of the fouling curve
where the fouling resistance increases progressively as depicted in Figure (1.3).

• Heat transfer coefficient


In single-phase forced convection, as the flow parameters bulk temperature and flow
velocity as well as the heat flux are kept constant, the overall heat transfer coefficient is
⎛ 1 (T − Tb ) ⎞
found to be inversely proportional to the surface temperature ⎜⎜ = s t ⎟⎟ . The term on
⎝U q& ⎠
the right hand side represents the overall resistance between the heat surface and the flow. In
a typical fouling run as time goes on, the accumulation of deposits on heat transfer surfaces
causes the heat transfer coefficient to decrease which in turn leads to an increase in the
surface temperature of the heat transfer surfaces just beneath the deposit layer.

3.3.4 Error analysis


Measured quantities are generally not exactly accurate since they are all subject to
systematic error (bias error) of the measuring devices and precision error which depends on
the fluctuation of a quantity at a single point of measurement (Kline and McClintock, 1953).
The combination of the bias error and the precision error is basically called uncertainty. In
this study, it is estimated to be with a 95% confidence as given in Eq. (3.11).

[
U = B2 + P2 ] 1
2
95% confidence (3.11)

It has been found that the largest uncertainty of fouling resistance occurs at the start of
all experiments when the temperature difference between heat transfer surface at time zero
and at the actual time is small. Contrariwise for the heat transfer coefficient, only a marginal
difference in uncertainty has been observed at the initial and final stages as shown in Figure
(3.9). This is somewhat expected since the driving temperature difference (Ts – Tb) is large in
both cases. Mathematical calculations of uncertainty are elaborated in detail in Appendix B.

44
3 Experimental Set-Up and Procedure

-3
1x10
Tb= 40 C; q&= 80 kW/m , v = 0.15 m/s; cb= 3.75 g/L, AISI 304 BA
o 2

3000

heat transfer coefficient [W/m K]


2
-4
8x10
fouling resistance [m K/W]
2500
-4
6x10
2

2000
initial stage

final stage
-4
4x10

1500
-4
2x10

1000

500
0 500 1000 1500 2000 2500
20
time [min.]

15
uncertainty [%]

fouling resistance
heat transfer coefficient
10

0
initial stage final stage

Fig. 3.9 Uncertainty of fouling resistance and heat transfer coefficient for a typical
fouling test (Tb = 40oC, v = 0.15 m/s, q& = 80 kW/m2 and cb= 3.75 g/L)

3.3.5 Matrix of operating conditions


A comprehensive set of experiments has been performed under forced convective heat
transfer fouling conditions for different modified surfaces. The range of operating conditions
is summarized in Table 3.2. The surface temperature was set to around 82oC in order to
comply with the surface temperature in the desalination unit of the MEDESOL project. In
addition, to define adequate experimental conditions where the influence of surface
characterizations can be explicitly seen, the fluid velocity of 0.15 m/s was set for all fouling
runs. In this case, the mass transfer control process has less influence than the reaction

45
3 Experimental Set-Up and Procedure

control process (Rizzo, 2008). For the grooved surfaces, a range of velocities from 0.1 to 0.3
m/s was adjusted for clean heat transfer runs.
Table 3.2 Range of operating conditions of the fouling experiments
Operating conditions Heat transfer surfaces
Parameter Unit Roughened Structured Modified
surfaces surfaces surfaces
o
bulk temperature C 40 40 40
velocity m/s 0.15 0.15 0.15
heat flux kW/m2 80-120 120 100-120
CaSO4 concentration g/L 3.75-4.4 4 4
o
surface temperature C 69-82 66-93 82

46
4 Experimental Results and Discussion

Chapter 4
EXPERIMENTAL RESULTS AND DISCUSSION

A total of 61 clean and fouling runs were performed during forced convective heat
transfer to investigate the precipitation fouling of CaSO4 solution on i) roughened; ii)
grooved; and iii) modified surfaces. Prior to each fouling run, the surface characteristics, i.e.
surface roughness, contact angle and surface energy were carefully measured. The
accumulation of deposits on modified surfaces has continuously been monitored by
measuring the variation in surface temperature with time. Both, fouling rate and induction
time were used to evaluate the influence of various modified surfaces on fouling mechanisms.
Subsequently, the experimental results were supported by post-fouling characterizations such
as visual examination of the deposit layer and examination of the deposit morphology.

4.1 Alteration of Surface Geometry


Surface irregularities in terms of surface roughness and surface periodic disturbance
in form of grooves constitute the majority of surface geometry modifications of this study
which endeavours to address the following points:
1. Impact of artificial surface roughness on surface wettability, fouling propensity
and adhesion.
2. Influence of the following geometrical parameters on fouling mechanisms:
• groove direction with respect to the direction of fluid flow
• groove dimensions in terms of width and depth

4.1.1 Fouling of roughened surfaces


4.1.1.1 Investigation of various roughened surfaces
Stainless steel surfaces AISI 304 BA with a wide range of surface roughness values
were investigated under constant operation conditions, i.e. bulk temperature and flow velocity
and different heat fluxes and bulk concentrations to understand the influence of these two
parameters on the fouling process. Sand papers with different grain sizes, i.e. P60, P80, P100
and P120 were used to provide a wide range of roughness. Here, "P" refers to the FEPA
(Federation of European Producers of Abrasives) standards. The roughness profile (Ra) of the
surfaces was measured by using the stylus instrument as illustrated in Figure 3.5 and its

47
4 Experimental Results and Discussion

values are listed in Tables 4.1. The values of Ra were selected to be between 0.6 to 1.6 μm.
Such choice of values was based on the following criteria:
i) a roughness profile between 0.02 μm to 0.6 μm will most likely not impact the
fouling process (Yoon and Lund, 1994; Santos et al., 2003),
ii) for roughness profiles above 2 μm, Zettler (2002) concluded that the effect of
increased turbulence and increased number of nucleation sites will be in balance.
In other words, no significant change in fouling propensity in terms of deposition
rate has been observed for surfaces having Ra values greater than 2 μm (Zettler,
2002).

Table 4.1 Investigated values of surface roughness


Specimens 1 2 3
Ra [μm] 0.54 (± 0.02) 1.1(± 0.05) 1.55 (± 0.12)

The effect of surface roughness on water contact angle is plotted in Figure 4.1. It is
apparent that increasing the surface roughness causes the contact angle to decrease. Based on
Eustathopoulos's hypothesis (Eustathopoulos et al., 1999) this trend can be attributed to the
effect of the sharp edge of the roughness profile. The sharp edge may pin the triple line
position far from a stable equilibrium state as shown in Figure 4.2. The triple line separates
solid (S), liquid (L) and vapour (V) surfaces.

90 a)

85 θo

80

75
θ [degree]

b)
70
θ h
65

60 dh
V
water contact angle
55 o
L
Τ = 20 C
50
S
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
Fig. 4.2 Equilibrium angle at
Ra [μm]
solid, liquid and vapour junction
Fig. 4.1 Influence of surface roughness on a) a smooth surface b) a rough
of stainless steel surfaces on the contact surface (Eustathopoulos et al.,
angle of water 1999)

48
4 Experimental Results and Discussion

Figure 4.3 presents the significant impact of variation of surface roughness on fouling
resistance as a function of time. As soon as nucleation starts, the fouling resistance is affected
significantly by the degree of roughness. Increasing surface roughness provides more zones
in terms of cavities or pits where flow velocities are very low (Rankin and Adamson, 1973).
This may encourage the deposition of initial crystals and consequently increase the build-up
of the fouling layer. More noticeably the fouling rate for the 0.54 μm surface is 1-2 times
lower than for the surface with a roughness of 1.55 μm.

-4
8x10
-6 2
q& = 80 kW/m2 [1] Ra= 0.54 μm, dRf/dt = 0.305x10 m K/W.min
-6 2
v = 0.15 m/s [2] Ra= 1.10 μm, dRf/dt = 0.502x10 m K/W.min
cb= 4 g/L -6 2
fouling resistance [m K/W]

-4 [3] Ra= 1.55 μm, dRf/dt = 0.955x10 m K/W.min


6x10 o
Tb= 40 C
2

-4
4x10

-4
2x10

0 200 400 600 800 1000


time [min.]

Fig. 4.3 Fouling resistance versus time for different roughened surfaces

Another plausible reason for the increase of fouling rate and fouling resistance can be
ascribed to the stronger adhesion on the rough surfaces which was clearly observed at the end
of each run. It was easier to remove the fouling layer from the smooth surface than from the
rough surfaces. Moreover, the visual examination of fouled surfaces indicates that the deposit
layer on the roughened surface with 1.55 μm was thicker than the deposit layer on the other
surfaces. The other difference that could be observed is the probability to have an
inhomogeneous deposit layer on the surface with 0.54 μm roughness. The inhomogeneity
basically refers to areas without deposits as can clearly be seen in Figure 4.4 (encircled
regions).

49
4 Experimental Results and Discussion

Ra = 0.54 μm Ra = 1.10 μm Ra = 1.55 μm

Fig. 4.4 Deposit layer on roughened surfaces for Tb = 40oC, Tso = 69oC, v = 0.15 m/s,
q& = 80 kW/m2 and cb= 4 g/L

4.1.1.2 Effect of contact angle on fouling mechanisms


A possible explanation for the effect of roughness on fouling mechanisms observed in
this study lies in the link between roughness, crystal nucleation and adhesion. Based on the
theory of nucleation proposed by Volmer (1929), the rate of heterogeneous nucleation ( J het )
can be estimated by an Arrhenius equation:

⎛ φ ΔGcrit
hom

J het

= C exp⎜ − ⎟ (4.1)
k BT ⎟
⎝ ⎠

where ΔGcrit
hom
represents the critical free energy barrier required to form a stable crystalline
nucleus. The free energy change that is accompanying a phase transformation is thought to
consist of two parts: one is related to the surface excess free energy, ΔGs, the excess free
energy between the surface of the particle and the bulk of the particle. The second part is
related to the volume excess free energy, ΔGv, the excess free energy between a very large
particle and the solute in the solution as shown in Figure 4.5. ΔG hom passes through a critical

maximum ΔGcrit
hom
corresponding to the critical nucleus size rcrit. The formation of critical

nuclei is favoured since ΔGcrit


hom
corresponds to a maximum energy decrease.

50
4 Experimental Results and Discussion

Fig. 4.5 Free energy diagram of homogeneous nucleation (Mullin, 1993)

The presence of foreign bodies such as the heat transfer surface causes the free energy
barrier to decrease, this case is called heterogeneous nucleation. The corresponding excess
free energy associated for heterogeneous nucleation is ( ΔGcrit
het
). It is always lower than

ΔGcrit
hom
for homogeneous nucleation. Thus,

ΔGcrit
het
= φ ΔGcrit
hom
(4.2)

where the nucleation correction factor φ is less than unity. Volmer (1939) provides an
equation to calculate φ , i.e.

φ=
(2 + cosθ )(1 − cosθ )2 (4.3)
4

Due to the nucleation correction factor φ , the volume of a critical nucleus (and thus
ΔGcrit) can be significantly smaller for heterogeneous nucleation, depending on the contact
angle. On the other hand, decreasing the contact angle causes the correction factor to
decrease, which possibly gives rise to the reduction of energy barrier (see Figure 4.6), and
consequently a higher fouling rate can be expected as clearly demonstrated in Figure 4.7.

51
4 Experimental Results and Discussion

hom
ΔGcrit

het
ΔG ΔGcrit

r
rcrit

Fig. 4.6 Difference in energy barrier for heterogeneous nucleation and homogeneous
nucleation

-5
1.4x10
q& = 120 kW/m2 cb= 4.4 g/L
-5 v = 0.15 m/s cb= 4 g/L
1.2x10 o
Ts= 82 C cb= 3.75 g/L
fouling rate [m K/W.min]

o
-5 Tb= 40 C
1.0x10

-6
2

8.0x10

-6
6.0x10

-6
4.0x10

-6
2.0x10

55 60 65 70 75 80 85

contact angle (θ) [degree]

Fig. 4.7 Effect of contact angle on fouling rate for various bulk concentrations

Another parameter that may correlate between surface roughness and fouling
propensity is the supersaturation ratio, “S” (ratio of bulk concentration to saturation
concentration of the solution). Figure 4.8 demonstrates that for a given contact angle
(roughness value) the fouling rate increases significantly with increasing supersaturation
ratio. For instance, increasing supersaturation ratio by 9.5% causes the fouling rate to
increase by approximately 40%. As the supersaturation ratio increases the chemical potential

52
4 Experimental Results and Discussion

( Δμ ) between the substance in solution (state 1, ' μ1 ') and in the crystal (state 2, ' μ 2 ') also
increases. Consequently, the critical free energy of the crystals will decrease as shown in Eqs.
(4.4) and (4.5). Thus, the number of crystals that have grown on the surface will increase with
increased supersaturation of the solution and consequently higher fouling rates are expected.
The predicted increase in the chemical potential is given in Table 4.2 for various
supersaturation ratios.

Δμ = RgTb ln S (4.4)

16πγ 3 (νR )
2
ΔG crit = (4.5)
3k B (Δμ )
2 2

Table 4.2 Chemical potential of calcium sulphate at Tb= 40oC


S Δμ [kJ/mol]
1.46 0.985
1.56 1.157
1.67 1.335
1.83 1.573

-6 q& = 120 kW/m2


8x10
v = 0.15 m/s
o
T s= 82 C
fouling rate [m K/W. min]

o
-6 T b= 40 C
6x10
2

-6
4x10

-6
2x10

R a=1.55 μm
0 R a=1.1 μm

1.4 1.5 1.6 1.7 1.8 1.9


supersaturation ratio, S

Fig. 4.8 Fouling rate as a function of supersaturation ratio for different roughened surfaces

53
4 Experimental Results and Discussion

It is also possible that the increase in concentration of CaSO4 causes the repulsive
1
electrostatic double layer energy to decrease due to a decrease in Debye length ( ) as shown
κ
1
in Figure 4.9. This is because the ionic strength dictates the value of which can be
κ
calculated using the following equation (Sharma and Rao, 2003).

1 0.304
= (4.6)
κ I

1 n
The ionic strength ( I = ∑
2 j
c j z 2j ) is closely linked to the bulk concentration

(Marshall et al., 1964). Increasing the bulk concentration causes the ionic strength to increase
and consequently thinner double layer thickness is expected. Such prospect may be a
plausible reason to reduce the energy barrier of adhesion which in turn leads to the
enhancement of adhesion between deposits and the surface. A similar effect of bulk
concentration on the Lifshitz-van der Waals and Lewis acid-base energies is not expected as
both of them are only marginally sensitive to the variation of bulk concentration.

0.78

0.76
Debye length, 1/κ [nm]

0.74

0.72

0.70
CaSO 4 solution
o
Tb = 40 C

0.68
0.15 0.16 0.17 0.18 0.19 0.20

ionic strength (I) [mol/l]

Fig. 4.9 Debye length as a function of ionic strength

54
4 Experimental Results and Discussion

Another parameter which is usually used to evaluate the fouling process is the
induction time (see Figure 1.3). As plotted in Figure 4.10, the induction time is inversely
proportional to surfaces roughness.

600
2
120 kW/ m
2
100 kW/ m
500
v = 0.15 m/s
cb = 4 g/l
400 o
Tb= 40 C
tind [min.]

300

200

100

0
0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

Ra [μm]

Fig. 4.10 Influence of surface roughness on induction time for various heat fluxes

This finding is consistent with the experimental results of Förster and Bohnet (2001).
Not surprisingly, the induction time decreases with increasing heat flux. This is because as
the heat flux increases, the initial surface temperature increases as well, resulting in a further
decrease in the solubility of the CaSO4 solution. This in turn increases the supersaturation of
salts with inverse solubility when the bulk concentration is kept constant and, therefore,
higher nucleation rate is expected. The reduction in induction time due to the increase of the
heat flux is 65-75% for both the lower and higher roughened surfaces.

55
4 Experimental Results and Discussion

4.1.2 Fouling of structured surface geometries


This section analyses the experimental fouling results of turbulent flow over
periodical grooves that are made on stainless steel surfaces. The categorisation of different
surface structures used in this study is shown in Figure 4.11. Grooves are classified into 1)
groove shapes in “V” form; 2) groove dimensions in terms of width and depth; and 3) groove
direction with respect to the direction of fluid flow, which are crossed, longitudinal and
mixed direction grooves. 2D and 3D sketches of “V” shaped grooves in terms of dimension
and direction are illustrated in Figures 4.12-4.14.

substrate material:
AISI 304 BA

structured geometry:
inclined grooves, ‘V’ shape, θ = 90o

groove dimensions

d = 0.1 mm d = 0.15 mm
w = 0.2 mm w = 0.30 mm

groove ctions

crossed grooves longitudinal grooves mixed grooves

Fig. 4.11 Flowchart of different grooved surfaces

Fig. 4.12 Crossed grooves (θ = 90o, Fig. 4.13 Longitudinal grooves (θ = 90o,
d = 0.1 mm and w = 0.2 mm) d = 0.1 mm and w = 0.2 mm)

56
4 Experimental Results and Discussion

Fig. 4.14 a) Mixed grooves with (d = 0.15 mm and w = 0.3 mm) and b) 3D view
of mixed grooves

4.1.2.1 Heat transfer over grooved surfaces


This section is written to provide a simplified idea of how the heat transfer occurs
within grooves. This may help in the analysis of the experimental results. Heat transfer within
grooves is more complex than that on flat plates. This is due to substantial turbulence caused
by the eddies generated within the channels which depend on the groove dimensions and their
direction to the flow. Therefore, before presenting heat transfer coefficients under both clean
and fouling conditions, it is indispensable to discern how the flow field changes as a result of
different surface structures and under which circumstances heat transfer surfaces may
perform better. As fluid flows over crossed grooves, eddies whose vortices are opposite to
those of the mainstream flow are generated that will cause an adverse pressure gradient.
Consequently, the generation of eddies in the groove itself may lead to strong mixing and
subsequently further agitation of the boundary layer. Another effect is due to the plume
emission at the corner of the grooves as shown in Figure 4.15. These plumes can be generated
in the shape of either backward or rotating eddies when there is abrupt change in the surface
texture as shown in this figure. Less eddies are expected for flow over mixed grooves and,
even more so, for longitudinal grooves. These hypotheses are consistent with the findings of
Chen et al. (2006), who showed that no additional pressure drag was generated in
longitudinal grooves while the crossed grooves caused stronger drag forces.

57
4 Experimental Results and Discussion

flow direction

plume

Fig. 4.15 Streamlines for a turbulent flow over crossed grooves

4.1.2.2 Clean heat transfer coefficient


Several experiments were carried out to determine the effect of flow velocity on the
clean heat transfer coefficient. As expected, as the velocity increases the heat transfer
coefficient also improves for all different groove geometries, regardless of their dimensions
and directions to flow. The main reason for this effect is the increasing turbulence level of the
bulk flow which in turn leads to a higher degree of turbulence within the grooves. Such
trends are clearly evident in Figure 4.16. A higher heat transfer coefficient was observed for
crossed grooves in comparison to longitudinal and mixed grooves, which both exhibit almost
the same heat transfer coefficients as velocity varies.
The groove dimensions (width and depth) also have a significant influence on the heat
transfer coefficient as shown in Figure 4.17. Increasing the groove width by only 0.1 mm
resulted in 10% higher heat transfer coefficients for lower velocities and up to 25% higher
values for higher velocities. This may be the result of more and larger eddies and
consequently higher level of mixing is expected in grooves with larger dimensions and vice
versa for smaller and narrower grooves. These findings are consistent with the results
reported by Wang (2003). He showed that the flow in narrow grooves is quite different from
that in wide grooves, since the streamlines do not produce eddies or plumes. Moreover, Wu et
al. (2002) confirmed that the Nusselt number increases with increasing groove depth. The
impact of groove dimensions is only marginal for the two other structures. Figure 4.18
depicts that while the flow velocity increases over longitudinal and mixed grooved surfaces
the differences between the two heat transfer coefficients are negligible, particularly for
lower flow velocities. This is a result of the generated streamlines around longitudinal and
mixed grooves, which do not show any resistance to the flow as this was the case for the

58
4 Experimental Results and Discussion

crossed grooves (Wu et al., 2002). For higher velocities, however, it can be concluded that
the magnitude of turbulence may depend more upon the surface structure, resulting in higher
heat transfer coefficients for the larger dimensions.

6.0
crossed grooves
heat transfer coefficient [kW/m K]

5.5 longitudinal grooves


mixed grooves
2

5.0 flat plate

4.5

4.0

3.5

2
3.0 q& = 120 kW / m
d = 0.15 mm
2.5 w = 0.3 mm
o
Tb= 40 C
2.0

0.10 0.15 0.20 0.25 0.30


flow velocity [m/s]
Fig. 4.16 Effect of flow velocity on the heat transfer coefficient of 'V' grooved surfaces

6.0
d = 0.15 mm, w = 0.3 mm
d = 0.10 mm, w = 0.2 mm
heat transfer coefficient [kW/m K]

5.5
2

5.0

4.5

4.0

3.5

3.0
q& =120 kW / m2
o
2.5 Tb= 40 C
crossed grooves
2.0
0.10 0.15 0.20 0.25 0.30
flow velocity [m/s]
Fig. 4.17 Variation of heat transfer coefficient with velocity over crossed grooves with
different dimensions

59
4 Experimental Results and Discussion

5.0
longitudinal grooves, w = 0.3 mm
longitudinal grooves, w = 0.2 mm

heat transfer coefficient [kW/m K]


4.5 mixed grooves, w = 0.3 mm
2
mixed grooves, w = 0.2 mm

4.0

3.5

3.0

2.5 q& =120 kW / m2


o
Tb=40 C
2.0
0.16 0.20 0.24 0.28 0.32
flow velocity [m/s]

Fig. 4.18 Heat transfer coefficient versus velocity for longitudinal and mixed grooves

4.1.2.3 Fouling of grooved surfaces

Surface structures with different shapes, dimensions and directions may also be of
interest when subjected to fouling. This is because, under fouling conditions, the generated
eddies and plumes, as shown in Figure 4.15, may give rise to substantially stronger drag
forces that may overcome the adhesion forces between the growing crystals and the heat
surface, and consequently lead to shearing-off of the early deposit. For longitudinal grooves,
a smaller flow resistance is anticipated and thus crystal nuclei are less likely to spall off the
surface. Figure 4.19 shows that, the induction time is a strong function of the groove direction
with respect to the fluid flow. Longer induction times occurred over the crossed grooves in
comparison with longitudinal and mixed grooves. As stated before, the crossed grooves
generate eddies whose vortices are opposite to those of the bulk flow. This would provide a
strong inhibition to the formation of initial nuclei.

Since a higher degree of turbulence is expected for the crossed grooves than for the
other groove directions (longitudinal and mixed), the build-up of the deposits on this surface
should be at a lower rate. This is plotted in relative form (Ut/Uo) in Figure 4.20. It can be seen
that a faster decrease in heat transfer performance occurred for mixed grooves in comparison
to that of crossed grooves. For instance, during the first 120 minutes of the experiment, the

60
4 Experimental Results and Discussion

maximum reduction in heat transfer performance was 58%, while in case of longitudinal
grooves it was 45%. However, such variation in heat transfer performance decreases with
increasing time. The stabilization in heat transfer coefficient occurred for all surfaces after
150 minutes of the experiment. A conceivable reason for such trend is that once the deposits
covered the surface entirely, the impact of groove direction has faded.

60
o
'V' grooved surface with θ =90
d = 0.1 mm
w = 0 .2 mm
50 crossed grooves

40
tind[m in.]

30

-2 longitudinal grooves
q& = 120 kWm
v = 0.15 m/s
20 mixed grooves
c b = 4 g/L
o
T b = 40 C

10
grooves direction

Fig. 4.19 Induction time as a function of groove directions

1.2
o 2
Tb = 40 C, q& = 120 kW/m , v = 0.15 m/s, cb = 4 g/L, d = 0.1 mm, w = 0.2 mm

1.0 crossed grooves


longitudinal grooves
mixed grooves
0.8
Ut / Uo

0.6

0.4

0.2

0.0
0 50 100 150
120 175
200
time [min.]

Fig. 4.20 Heat transfer performance as a function of groove directions

61
4 Experimental Results and Discussion

Figure 4.21 presents three pictures of deposit layers on different surface structures for
the same dimensions and operating conditions.

a) crossed grooves b) longitudinal grooves c) mixed grooves

Fig. 4.21 Fouling layer on 'V' grooved surfaces with d = 0.1 mm and w = 0.2 mm for
Tb = 40oC, v = 0.15 m/s, q& = 120 kW/m2 and cb= 4 g/L

Visual examination of these fouled surfaces after each run indicates that the crossed
groove surface has a fragile and thin deposit layer, while for longitudinal and mixed grooves
the fouling layer was relatively thick and adhesive. The impact of entrapped eddies is also
evident inside the crossed grooves, in which either no or only little deposit was observed, see
Figure 4.21-a (enclosed area). Contrariwise, a homogenous deposit layer is evident for the
other two structures.

4.1.2.4 Effect of groove dimensions on fouling mechanisms


Increasing the groove dimensions overall has a positive impact on fouling propensity,
particularly for the crossed groove surface as depicted in Figure 4.22. Increasing the depth of
grooves led to a comparatively higher reduction in fouling rate. No such trend was seen for
longitudinal and mixed grooves where both dimensions exhibit the same fouling rate. The
reduction in fouling rate can be attributed to a high level of mixing within the grooves due to
the entrapped eddies inside larger cavities. The same eddies may sweep away initial crystal
nuclei that attempt to form inside the grooves. However, after a certain period of time the
fouling curves level off abruptly (see section A-Figure 4.22). Similar behaviour was also
reported by other studies (Santos et al., 2003; Augustin et al., 2007). It can be speculated that
the sharp reduction of fouling resistance might have been due to sub-cooled flow boiling

62
4 Experimental Results and Discussion

effects. In other words, for a porous deposit layer as in the case of grooved surfaces the liquid
may permeate the pores and probably arrives at the deposit/solid interface where the surface
temperature exceeds the boiling temperature. This may provide conditions for bubbles to be
generated within the pores of the deposit layer, which in turn prevents the further build-up of
deposit. As a consequence, a sharp decrease in deposition rate may occur. A similar
phenomenon has also been observed during the investigation of water based coated surfaces
(WB) as can be seen in Figure 4.27. However, after an experimental duration of around 250
minutes the deposition process overcame the removal process leading to a second increase of
fouling resistance but at a lower rate in comparison to that of the initial stage.

-3
1.4x10
1 52 section A -2
q& = 120 kWm
1.2x10
-3 1 50
v = 0.15 m/s
Ts [ C ]

cb = 4 g/L
o

1 48
fouling resistance [mK/W]

o
-3 1 46
Tb = 40 C
1.0x10
1 44
crossed grooves
2

-4 1 45 150 1 55 1 60 16 5
8.0x10 tim e [m in.]

6.0x10
-4
A

-4 -6 2
4.0x10 1 [1] dRf/dt = 9.19x10 m K/W min
-6 2
[2] dRf/dt = 6.25x10 m K/W min
-4
2.0x10 2

[1] d = 0.10 mm, w = 0.2 mm


0.0
[2] d = 0.15 mm, w = 0.3 mm

0 100 200 300 400 500 600 700


time [min.]

Fig. 4.22 Fouling resistance versus time for crossed grooves with variable dimensions

63
4 Experimental Results and Discussion

4.2 Alteration of the Surface Energy Related Properties


As explained in Chapter 2, alteration of the surface energy of heated surfaces is
considered as a novel anti-fouling technique. To scrutinize possible variation, the stainless
steel surfaces (AISI 304 BA) were modified by the following coating materials:
• solvent and water based (SB and WB) coatings
• electroless Ni-P coatings
• nano-structured coatings
The experimental work of this part followed three main steps:
1) Characterization of modified surfaces.
2) Fouling behaviour on modified surfaces.
3) Analysis of the fouling results which mainly consisted of the examination of; i) the
impact of surface energy components on the CaSO4 deposition process, ii) CaSO4
deposit morphology, and iii) reproducibility of fouling experiments of modified
surfaces.

4.2.1 Surface characterizations


4.2.1.1 Coating composition and preparation
• Solvent based coatings (SB1-SB4): They are prepared by directly dispersing a
mixture of various materials (hexagonal boron nitride (BN), aluminium silicate as
filler, fluorinated polyethylene wax) in the butylacetate diluted binder (Silicone
polyester resin) solution. The main difference between the different SB coatings is
the percentage of wax in the coating composites.

• Water based coating (WB): It is produced by directly dispersing the components


(hexagonal boron nitride (BN), titanium dioxide (TiO2) as filler, fluorinated
polyethylene wax) in water.

SB and WB coatings are supplied by ItN Nanovation AG, Saarbrücken, Germany.


The coating process follows this procedure: firstly the stainless steel surfaces were cleaned
and degreased to remove contaminations before the coating takes place. Afterwards, various
solvent and water based solutions were used for dissolving the ingredients of the coating
materials. In both cases, the material was laid on as a wet layer with thickness of
approximately 100 μm by a low pressure spray gun. The surfaces were then ventilated under

64
4 Experimental Results and Discussion

room temperature for 15 minutes. The coatings were then consolidated by temperature
annealing in order to harden and compress the coating layers to about 50-55 μm for all
surfaces. The WB coating was hardened for 3 hours at 120oC, while only 30 minutes were
enough to harden the SB coatings at 230oC.

• Electroless Ni-P coatings: These coatings are supplied by NovoPlan GmbH,


Aalen, Germany. The first type of coating consists of 50 µm of Ni-P with 7-9% P
(phosphor) while the second coating which is called Ni-P-BN consists of two
layers. The first layer consists of 40 µm of Ni-P with 7−9% P and the second layer
on top is made from 10 µm of BN (Ni−P with 12-15% boron nitride). This is
because the present technology of the coating company is limited to a maximum
coating thickness of "only Ni−P−BN" of 15 µm. Thus, an additional layer of Ni−P
is applied below the second layer in order to realize an identical thickness of 50
µm for both coatings of Ni−P and Ni−P−BN.

• Nano-structured coatings (K-S1 and K-S2): These coatings are supplied by


Interlotus Nanotechnologie GmbH in Germany. The coatings include silicon
nano-particles that are spread on the substrate to increase hydrophobicity: The
thicknesses of the coatings are 1.5 and 2.5 μm, respectively.

4.2.1.2 Physical surface properties


The values of surface roughness for all modified surfaces are depicted in Figure 4.23
in three groups of surface roughness. The first group, mainly consisting of SB coatings, has
high surface roughness values in a range of 1-1.3 μm. The remaining groups possess a surface
roughness in a range of 0.5-0.8 μm and 0.1-0.3 μm for electroless Ni-P and nano-structured
coated surfaces. Table 4.3 lists the values of water contact angle and surface energy
components for the modified surfaces. The SB, WB and the nano-structured coated surfaces
have a large water contact angle. A minor change has been observed for the Ni-P coated
surfaces with and without boron nitride. Interestingly, K-S1 and SB2 coated surfaces are the
only ones that are characterized as hydrophobic with a contact angle of greater than 90o.

65
4 Experimental Results and Discussion

1,6

group 1
1,4

1,2

1,0
group 2
Ra [μm]

0,8

0,6

0,4 group 3

0,2

0,0
SB1

SB3

WB
SB2

SB4

K-S1

K-S2
Ni-P-BN

Ni-P
coating

Fig. 4.23 Surface roughness of modified surfaces

Table 4.3 Water contact angle and surface energy components for untreated and modified
surfaces
Type of surface Water Surface energy components, mJ/m2
contact
angle, γ2 γ 2LW γ 2+ γ 2−
degree
untreated stainless AISI 304 73.3 35.76 35.11 0.03 3.46
steel surface BA
solvent/water based SB1 85.2 24.32 23.73 0.01 8.59
coated surfaces SB2 92.2 18.81 18.57 0.001 13.85
SB3 87.4 27.63 27.36 0.0021 9.26
SB4 86.5 25.57 24.3 0.05 8.17
WB 89.1 30.17 27.75 0.1 14.65
electroless Ni-P Ni-P 75.1 48.15 40.68 0.84 16.64
coated surfaces Ni-P-BN 62 52.29 45.55 0.45 25.3
nano-structured coated K-S1 94 21.80 21.36 0.003 16.55
surfaces K-S2 82.9 25.04 23.17 0.09 9.74

As can be seen in Figure 4.24, SB and nano-structured coatings reduce γ 2 by


approximately (23.5% - 47.4%) in comparison to that of untreated stainless steel surfaces,
while for the WB coating this is only 15.6%. It should be pointed out that the electroless Ni-P
coatings with and without boron nitride show significantly larger γ 2 values. Furthermore the

66
4 Experimental Results and Discussion

Lifshitz-van der Waals surface energy component ( γ 2LW ) has largely been affected by the

type of modification as illustrated in Table 4.3. The SB and WB coatings reduce γ 2LW by
approximately 20-46% in comparison to the stainless steel surface. Similar reduction has also
been achieved by using the nano-structured coatings, while in case of Ni-P coatings a
significant increase of more than 15% has been observed.

electroless Ni-P coated surfaces


Ni-P-BN

Ni-P

K-S2 nano-structured coated surfaces


K-S1
surface

WB water based coated surface

SB4

SB3

SB2

SB1 solvent based coated surfaces

AISI 304 BA untreated surface

0 5 10 15 20 25 30 35 40 45 50 55
2
γ2 [mJ/m ]

Fig. 4.24 Total surface energy of modified and untreated surfaces

However, the most distinguishable factor that changes from one surface to another is
the electron donor component ( γ 2− ) (see Table 4.3). For all modified surfaces γ 2− is higher
than the value of the untreated stainless steel surface. For instance, in case of Ni-P and Ni-P-
BN, γ 2− is approximately 5-7 times higher than the value of the stainless steel surface. A

higher values of γ 2− has also been observed for the nano-structured (K-S1) coating. On the

other hand, SB1 and SB4 coated surfaces exhibit approximately similar γ 2− as both coatings
have the same qualitative and quantitative material compositions except the wax content.

67
4 Experimental Results and Discussion

4.2.1.3 Resistance of coatings to abrasion and corrosion


The mechanical and chemical stabilities of coatings are the main technical hurdles
that may restrict their utilization as anti-fouling surfaces. Abrasion and corrosion tests are
common tests which are usually carried out to determine such stabilities. In this study only
the SB and WB coated surfaces have been examined at ItN-nanovation AG, Saarbrücken.
The company has the facilities to do such tests in cooperation with ITW.

• Abrasion resistance
In this study, abrasion was examined by the resistance to peeling off from the surface.
The scope of the abrasion test ranges from estimation of the lifetime performance of coatings
to actual quantification of mechanical and physical degradation. A short lifetime of the
coatings may cause additional adverse effects such as prompting faster nucleation on the
damaged spots which in turn leads to sharp decrease in heat transfer efficiency of the heat
exchanger. The abrasion test involved a defined quantity of sand (about 3.5 kg) which was
poured from a height of about 1.5 meters at an angle of 45° upon the coated surfaces under
room temperature. According to DIN 52348, the loss in coating thickness is measured against
the weight, which is called poured sand coefficient. The test is terminated as soon as the
coating at one point is completely wiped off from the surface as shown in Figure 4.25. The
results show that the SB coatings are much more stable (about 10 folds) to abrasion than the
WB coating.

a) before abrasion test b) after abrasion test

damaged area of
the coating

Fig. 4.25 Abrasion test of SB coated surface

• Corrosion resistance
Since the coating layer is desired to exercise maximum resistance to corrosion
through the intended lifetime of the modified surface in use, comprehensive corrosion tests

68
4 Experimental Results and Discussion

were carried out to asses the corrosion-resistivity of the coatings. Among various such tests,
the salt test provides a severe corrosion environment which exerts a corrosive attack on the
coated surfaces. 20% NaCl salt solution was applied onto the coated surface at 120oC for 2
hours. Afterwards, the dried salt was carefully shaked and rinsed from the surface. Figure
4.26 shows that the salt crystals are completely swept off after rinsing with water. A similar
effect has been observed for NaCl/CaCl2 and CaCl2/CaSO4 solutions. These laboratory results
show that the SB and WB coatings have a good resistance to corrosion.

after 2 hours contact with 20% after shaking after rinsing


NaCl solution at 120oC

Fig. 4.26 Result of the corrosion test for the solvent based (SB) coating

4.2.2 Fouling investigation of modified surfaces


4.2.2.1 Solvent/water based coated surfaces
From the data shown in Figures 4.27 and 4.28, one can immediately see that SB and
WB coated surfaces reduce the fouling resistance significantly in comparison to the untreated
stainless steel surface. For instance, during the first 5 hours the average reduction in fouling
resistance for SB2, SB3, SB4 and WB coated surfaces were 66%, 8%, 6% and 13%,
respectively and almost 100% for SB1. SB and WB coated surfaces initially behave like the
untreated stainless steel surface except for the SB2 surface. The important point here,
however, is not only the quantitative results but also the qualitative information that is
depicted in Figures 4.27 and 4.28. For the SB coated surfaces (except in the case of SB2) the
deposits tended to spall-off the surface after a short period of time, followed by a second
deposition process. Such behaviour has not been observed for the WB coated surface.
Nevertheless, in most cases of SB coated surfaces, the second deposition process started
almost from zero fouling resistance and at lower fouling rate. The fouling behaviour for the
SB4 coated surface differs somewhat from that of the other coated surfaces. After
approximately 600 minutes of experiment, only part of its deposit layer spalled off the
surface and the remaining deposits started to grow at a very low fouling rate in comparison to
that of the initial stage.

69
4 Experimental Results and Discussion

-3
1x10
o
Tb= 40 C [1] stainless steel
v = 0.15 m/s [2] WB
fouling resistance [m K/W] 8x10
-4 [3] SB1
cb= 4 g/L
[4] SB2
2

6x10
-4 1
2

-4
4x10
3

-4
2x10
4
0

0 200 400 600 800 1000 1200 1400

time [min.]
Fig. 4.27 Fouling resistance as a function of time for SB, WB and untreated surfaces

-3
1x10
Tb= 40 C
o [1] stainless steel
[2] SB3
1x10
-3 v = 0.15 m/s [3] SB4
cb= 4 g/L
fouling resistance [m K/W]

8x10
-4
3
2

-4
6x10
1
-4
4x10

-4
2x10
2
0

0 400 800 1200 1600 2000


time [min.]

Fig. 4.28 Fouling resistance as a function of time for SB3, SB4 and untreated surfaces

4.2.2.2 Electroless Ni-P coated surfaces


Figure 4.29 illustrates the variation of fouling resistance with time for the electroless
Ni-P and Ni-P-BN coated surfaces and for the untreated surface. Similar to SB and WB

70
4 Experimental Results and Discussion

coatings, a significant reduction in fouling resistance has generally been observed. Initially
the Ni-P coated surface behaves like the untreated surface, while the Ni-P-BN coated surface
remains free from any deposit during the first 5 hours. It is evident that the Ni-P-BN coated
surface has a noticeably different fouling behaviour than the Ni-P and untreated surfaces,
since longer induction time and lower fouling rate are the main characteristics of the Ni-P-BN
coated surface. Furthermore, after a certain period of time the CaSO4 deposits on Ni-P-BN
surface tended to spall off the surface, followed by a second deposition process at lower
progressive rate (lower fouling rate).

-4
8x10
[1] stainless steel
[2] Ni-P
[3] Ni-P-BN
fouling resistance [m KW]

-4
6x10
2

1
-4
4x10

2
-4
2x10
v = 0.15 m/s
3
cb= 4 g/L
0 o
Tb= 40 C

0 400 800 1200 1600

time [min.]

Fig. 4.29 Fouling resistance as a function of time for the untreated and the electroless Ni-P
coated surfaces

4.2.2.3 Nano-structured coated surfaces


The influence of the nano-structured coated surfaces (K-S1 and K-S2) on fouling
resistance is illustrated in Figure 4.30. Initially K-S1 behaves like the untreated surface, while
the K-S2 coated surface remains free from any deposit during the first 5 hours. In other
words, the fouling resistance for the K-S1 coated surface was in average reduced to less than
40% in comparison to the stainless steel surface. In case of K-S2, nevertheless, 100%
reduction in fouling resistance has been observed in this time span. It is evident that the K-S2
coating increases the duration of the induction time as much as 9-fold over that of the

71
4 Experimental Results and Discussion

untreated stainless steel surface, followed by an initial fouling rate which is about 38% lower
than that of the stainless steel surface. As time goes on, a significant difference in fouling
behaviour was observed. After a duration of approximately 700 minutes, part of the deposits
spalled off the surface, followed by a second deposition process with a fouling rate
approximately similar to that of the initial stage. However, after 800 minutes the fouling
resistance decreased rapidly, reaching a low fouling resistance value (1.15x10-4 m2K/W).

o
T b= 40 C [1] stainless steel
-4
8x10 [2] K-S1
v = 0.15 m/s
[3] K-S2
c b= 4 g/L
fouling resistance [m K/W]

-4
6x10 1
2

2
-4
4x10

-4
2x10
3

0 200 400 600 800 1000 1200


time [min.]

Fig. 4.30 Fouling resistance as a function of time for the nano-structured and untreated
surfaces

Afterwards, a third deposition process occurred with 80% reduction in fouling rate in
comparison to that of the initial fouling rate, reaching an asymptotic value after
approximately 17 hours. A similar trend was observed for the K-S1 coated surface. After the
detachment process, the fouling resistance decreased rapidly, reaching a negative value for
the fouling resistance, i.e. enhancement of heat transfer. As time goes on, the amount of
deposit seems to be large enough to outweigh the increase in heat transfer coefficient and the
fouling resistance increased again until a second detachment off process occurred. At this
stage, the deposits gradually detached the surface until they reached a constant fouling
resistance. At the end of the run, the final fouling resistance for the K-S2 coated surface is
less than that for the K-S1 surface by approximately 50%.

72
4 Experimental Results and Discussion

The comparison of fouling resistances of all modified surfaces and the untreated
stainless steel surface highlights the following important features:

o Reduction of fouling resistance


Figure 4.31 summarizes the percentage reduction of fouling resistance and
heat transfer performance (Ut/Uo) for all of the modified surfaces after the first 5
hours of the experiment. The comparison provides a tool to asses the impact of the
coated surfaces on fouling propensity. As can be seen, altering the surface properties
of the stainless steel surfaces causes in most cases a significant decrease in fouling
resistance and, as a consequence, a lower reduction in heat transfer performance.

8
fouling resistance
6
Rf [10 m K/W]

4
2
-4

00
120
reduction of fouling resistance
(|Rf,co- Rf,ss|/Rf,ss)x100%

100

80

60

40

20

1.20
heat transfer performance
1.0

0.8
Ut / Uo

0.6

0.4

0.2

0.0
AISI 304 BA

K-S1
SB1

Ni-P
SB3

SB4

K-S2
Ni-P-BN
SB2

WB

surface

Fig. 4.31 Reduction of fouling resistance and heat transfer performance for untreated and
modified surfaces at Tb = 40oC, v = 0.15 m/s and cb = 4 g/L at t = 5 hours

73
4 Experimental Results and Discussion

Contrariwise, the SB3, SB4, WB and K-S1 coatings cause less than 40%
reduction in fouling resistance in comparison to that of the stainless steel surface. This
is analogous to a reduction of heat transfer performance by approximately 50-60% in
comparison to the value at clean condition. However, for the cumulative verification
of fouling results and for the examination of the previous models, the initial fouling
rate and induction time are used instead of the fouling resistance.

• Spalling off
Once a certain deposit thickness has been reached, the shear force may be
higher than the adhesion bonds between the CaSO4 deposits and the surface, leading
to detachment of part of the deposit layer. It may further be speculated that the
remaining lumps of deposits will increase the level of turbulence close to the heat
transfer surface. The presence of this additional turbulence together with the existence
of driving forces due to temperature gradient and concentration may lead to
subsequent (secondary) fouling rates, which are lower than the initial fouling rates as
shown in Figure 4.32, except for the case of Ni-P-BN where its secondary fouling rate
is approximately 28% higher than the initial fouling rate.

-6
5x10
initial fouling rate
secondary fouling rate
-6
4x10
fouling rate [m K/W.min]

-6
3x10
2

-6
2x10

-6
1x10

0
SB3 SB4 K-S1 K-S2 Ni-P-BN
coating
Fig. 4.32 Initial and secondary fouling rates for modified surfaces

• Limited impact of surface roughness

74
4 Experimental Results and Discussion

As described in section 4.1, if the surface roughness of stainless steel surfaces


increases, then more nucleation sites for deposition are available giving rise to higher
fouling rates. Contrariwise, no direct relationship between the surface roughness of
modified surfaces and fouling rates has been observed as depicted in Figure 4.33.
Such differences may be due to the influence of surface energy properties on the
fouling mechanisms. On the other hand, the dominant influence of surface energy
over surface roughness on fouling propensity may cause the impact of the latter to be
very limited. This finding is consistent with that of Beuf et al. (2003) who showed
that the free surface energy plays a predominant role and the roughness a minor role
with respect to the level of fouling and cleaning efficiency of milk product during
convective heat transfer. However, such speculation requires further investigation to
verify what will happen if the same coating is applied on different roughened
surfaces.

-6
5x10

Ni-P AISI 304 BA


SB1
initial fouling rate [m K/W. min]

-6
4x10 WB
SB3 SB4
2

-6
3x10
K-S1
K-S2
-6
2x10

SB2
-6
Ni-P-BN
1x10

0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4


Ra [μm]

Fig. 4.33 Fouling rate as a function of surface roughness for the


modified and untreated surfaces

4.2.3 Post-fouling analysis of the deposit layer on modified surfaces


• Visual examination of the deposit layer
After termination of each experiment, the specimens were carefully dismantled from
the test rig and then from the specimen holder. Great care has been taken to avoid any
damage to the deposit layer. Afterwards, they were visually inspected to discern the influence

75
4 Experimental Results and Discussion

of coatings on the CaSO4 deposit layer. Figure 4.34 presents some pictures of the deposit
layers on the coated surfaces as well as on the untreated stainless steel surface for similar
operating conditions. In case of SB coated surfaces, the deposit layer was relatively thin and
non-homogenous, as parts of the surfaces were free of any deposit. For the WB coated
surface the deposit layer was relatively thick, homogenous and approximately similar to that
of the stainless steel surface. However, the WB coated surface exhibited finer and more
fragile crystals in comparison to that of the stainless steel deposit layer.
For the SB coated surfaces, the deposit layer also spalled off as clearly shown in
Figure 4.34. It can be seen that i) no deposit layer or even patchy crystals can be found on the
SB1 coated surface, ii) patchy crystals with weak adhesion have been found on the SB2
coated surface, and iii) porous deposit layers have been observed for SB2 and SB4 coated
surfaces. The presence of pore spaces in the deposit layer prevents the individual crystals
from getting together to form a hard lattice structure. As a consequence, the deposits adhered
more loosely and could easily be washed away from SB coated surfaces. Based on these
observations, the SB coated surfaces provide much better non-stick conditions than the WB
and stainless steel surfaces.
Unlike SB coated surfaces, the deposit layers on Ni-P coated surfaces were relatively
thick and homogenous. Except for Ni-P-BN, for which the deposit layer was not homogenous
(see enclosed area, Figure 4.34) and seems to be porous compared to stainless steel and Ni-P
coated surfaces. The visual examination of the deposit layers on nano-structured (K-S1 and
K-S2) surfaces indicates that the deposit layer on K-S1 coated surface was very thin with
better non-stickability, as the deposit layer fell off during the dismantling of the surface from
the test section. It is noteworthy that a change in colour occurred only for the K-S1 coated
surface. For the K-S2 coated surface only few deposits have been observed which could be
easier removed than the tenacious deposit layer on the stainless steel surface. The most
noticeable difference was for K-S2 that some spots of oxidization were observed (encircled
regions) which could be a result of poor abrasion resistance of the coating.

76
4 Experimental Results and Discussion

AISI 304 BA SB1 SB2 SB3 SB4

WB K-S1 K-S2 Ni-P Ni-P-BN


77

changed color area

Fig. 4.34 Deposit layer on coated and untreated stainless steel surfaces at Tb = 40oC, v = 0.15 m/s, q& = 100 kW/m2 and cb= 4 g/L

77
4 Experimental Results and Discussion

• Deposit morphology
Post-fouling deposit morphology analysis can provide some useful information about
the crystal structure. Since the initial surface temperature was always lower than 97oC, only
calcium sulphate dihydrate is expected to be deposited on the heating surface (Glater et al.,
1980). The formation of calcium sulphate anhydrite can be assumed to be negligible as it is a
very slow process (Hasson and Zahavi, 1970). This assumption has been confirmed by an
elemental analysis of deposits at the Department of Chemistry, the University of Stuttgart.
Since the deposit layer is expected to be mainly in dihydrate phase, a needle-shaped
structure of crystals would be expected. This is clearly shown in the SEM pictures in Figure
4.35. The pictures were taken using a SEM (MIRA-II LMH-IB, with a magnification range of
up to 1,000,000 X) for a 4 g/L salt concentration and a heat flux of 100 kW/m2. To obtain
such pictures, samples of deposit have been taken at the end of the experiments with great
care to avoid any damage to the deposit structure. It was undoubtedly found that the type of
coating significantly influences the structure of the calcium sulphate deposit. The differences,
in terms of size and orientation, are clearly noticeable. In particular, larger, thicker and
horizontally orientated crystals are the main characteristics of the WB and Ni-P coated
surfaces (see Figures 4.35c and 4.35d), while smaller, longer and randomly orientated
crystals are typical for SB and Ni-P-BN coated surfaces as shown in Figures 4.35a, 4.35b,
and 4.35e. In general, the crystals on WB coated surface were harder and more adherent than
those on SB coated surfaces. In contrast, the crystals on the SB and Ni-P-BN coated surfaces
were loosely attached to the surface as they could easily be washed away by just scrubbing
the surface with HCl diluted solution. It should also be pointed out that for surfaces having
longer induction time such as Ni-P-BN, the crystals are characterized by long needle-shaped
crystals. Unfortunately, it was difficult to obtain deposit samples for the nano-structured
coated surfaces (K-S1 and K-S2) since, as mentioned before, the deposit layer for the K-S1
coated surface fell-off during dismantling of the surface from the test section, while for the
K-S2 coated surface the fouling layer consisted only of few deposits.

78
4 Experimental Results and Discussion

SB3 SB4

(a)
(b)
WB Ni-P

(c) (d)

Ni-P-BN

(e)

Fig. 4.35 SEM images of the CaSO4 deposit morphology for the modified surfaces (size of
SEM pictures is 500 μm)

4.2.4 Influence of surface energy on fouling propensity


The preceding sections described the impact of the surface modifications on fouling
resistance and deposit layer without referring to the factors affecting fouling mechanisms. As
mentioned in Chapter 1, the fouling process during convective heat transfer depends mainly
on the physicochemical properties of the foulant, operating conditions and surface
characteristics. As the first two groups (concentration, bulk temperature and flow velocity)

79
4 Experimental Results and Discussion

are kept constant for the investigation of modified surfaces, then the only remaining dominant
fouling parameters are the surface-related properties such as interfacial surface energies
between deposits and the modified surfaces. For better and cumulative comparison, Table 4.4
lists the induction time and fouling rate for all investigated modified surfaces.

Table 4.4 Fouling rate and induction time for untreated and modified surfaces
Type of surface Initial fouling rate Induction time
[x 10-6 m2 K/W.min] [min]
untreated stainless steel AISI 304 BA 4.31 43
surface
solvent/water based coated SB1 4.1 89
surfaces SB2 1.171 181
SB3 3.9 58
SB4 3.93 72
WB 4.22 85
electroless Ni-P coated Ni-P 4.31 105
surfaces Ni-P-BN 0.961 580
nano-structured coated K-S1 2.96 130
surfaces K-S2 2.66 450

In Figures 4.36 and 4.37 the initial fouling rate and the induction time are plotted
against the total surface energy ( γ 2 ). It is evident that, the modified surfaces with low surface
energy (like SB1) did not show any significant reduction of fouling rate in comparison to that
of the untreated stainless steel surface. For Ni-P-BN, however, a reduction of 75% in fouling
rate was found even though its surface energy is 46% larger than the stainless steel. In
addition, the deposition of calcium sulphate started after different induction times. In most
cases, reducing the total surface energy causes a marginal increase in induction time.
However, in case of the K-S2 coated surface a longer induction time has been observed. It
was 9-10 times higher than for stainless steel. The increase of induction time for K-S2 is
probably due to the deterioration of the coating layer that could detach the initial deposits
from the surface. This can be confirmed by inspection of Figure 4.34, from which it is
evident that a large portion of the coating layer is damaged and only little deposits remain on
the surface. Hence the removal of the deposit is not solely due the effect of surface energy,
but the coating itself may have taken away part of the deposit while getting detached from the
substrate. This may account for the significant reduction in fouling resistance and fouling rate
as shown in Figure 4.36. The behaviour of the K-S2 coated surface seems somewhat similar
to that of the Ni-P-PTFE coated surface investigated by Zettler (2002). He showed that part

80
4 Experimental Results and Discussion

of the significant reduction or the almost complete elimination of fouling was due to the
deterioration of the coating layer.

5.0
initial fouling rate [10 m K/W.min]

4.5 AISI 304 BA


SB1 WB Ni-P
4.0
SB4
2

3.5
SB3
-6

K-S1
3.0
K-S2
2.5

2.0

1.5
SB2
1.0 Ni-P-BN

0.5
15 20 25 30 35 40 45 50 55
2
γ2 [mJ/m ]

Fig. 4.36 Initial fouling rate as a function of the total surface energy for untreated and
modified surfaces

600
Ni-P-BN

500
K-S2
400
tind [min.]

300

SB2
200

K-S1
100 SB4 Ni-P
SB1 WB
AISI 304 BA
SB3
0

15 20 25 30 35 40 45 50 55
2
γ2 [mJ/m ]

Fig. 4.37 Induction time as a function of the total surface energy for untreated and modified
surfaces

81
4 Experimental Results and Discussion

Based on the previous observations, it can therefore be concluded that the total
surface energy alone is not sufficient to describe the deposition process. These findings are
consistent with those of Augustin et al. (2005) and Geddert et al. (2009) who showed that the
comparison of the total surface energy of DLC-coatings customized with variable build-in of
additional elements such as SICAN (a:C:H:Si) and SICON (a:C:H:Si:O) yields no clear
tendency for the prediction of fouling behaviour. In contrast, Zhao et al. (2005a) found that
the adhesion of both bacteria and CaSO4 deposit was minimal when the total surface energy
was in the range of 26-30 mJ/m2.
To examine the influence of other components of surface energy, Figure 4.38 shows
the relation between γ 2LW and fouling rate for all modified surfaces. A quick examination of

this figure shows that no general correlation between γ 2LW and fouling rate can be

established. In addition, a comparison between γ 2LW and induction time for all modified
surfaces as shown in Figure 4.39 also does not provide direct correlation. This implies that
γ 2LW would not help to interpret fouling results which is consistent with similar comparisons
reported by Geddert et al. (2009). Thus other surface properties should be explored that
would influence fouling in more systematic manner.

5.0
initial fouling rate [10 m K/W. min]

AISI 304 BA
4.5
SB1 WB Ni-P
4.0
SB3 SB4
2

3.5
-6

K-S1
3.0

2.5 K-S2

2.0

1.5
SB2
1.0 Ni-P-BN

0.5
15 20 25 30 35 40 45 50 55
LW 2
γ2 [mJ/m ]

Fig. 4.38 Variation of initial fouling rate with Lifshitz-van der Waals component of untreated
and modified surfaces

82
4 Experimental Results and Discussion

600
Ni-P-BN

500
K-S2
400
tind [min.]

300
SB2
200

K-S1 SB1 WB Ni-P


100
SB4
0 SB3 AISI 304 BA

15 20 25 30 35 40 45 50
LW 2
γ2 [mJ/m ]

Fig. 4.39 Induction time as function of Lifshitz-van der Waals component for untreated and
modified surfaces

As mentioned in section 4.2.1.2, the electron donor component ( γ 2− ) can potentially

be used to correlate fouling propensity. Indeed, this is due to the significant variation of γ 2−
by surface modification. Figure 4.40 demonstrates how the fouling rate varies as a function of
γ 2− . It is evident that the higher γ 2− , the lower is the fouling rate, as in the case of Ni-P-BN
coating where its fouling rate is 77% lower than that of the untreated stainless steel surface.
As explained in Chapter 2, four types of interaction energies affect the interfacial interaction
between a deposit and a heated surface. One of them is the Lewis acid-base interaction
energy. A closer look at Eq. (2.13) shows that increasing γ 2− seems to be the prominent factor
that causes higher values of the Lewis acid-base energy to be more repulsive than attractive
in comparison to the stainless steel surface. Therefore, lower fouling rates are expected for
surfaces having higher electron donor components. These findings are consistent with those
of Visser (2001) who found that an important parameter to reduce calcium phosphate fouling
is to increase the electron donor component of the surface energy.

83
4 Experimental Results and Discussion

5.0

initial fouling rate [10 m K/W.min]


AISI 304 BA
4.5
Ni-P
SB1
4.0 WB
SB4
2
-6
3.5
SB3

3.0 K-S1
K-S2
2.5

2.0

1.5
SB2 Ni-P-BN
1.0

0.5
0 5 10 15 20 25 30
- 2
γ2 [mJ/m ]

Fig. 4.40 Initial fouling rate as a function of electron donor component for untreated and
modified surfaces

There are, however, several investigations which reported an opposite trend, i.e. the
deposition rate of calcium phosphate is increased for higher values of γ 2− (Wu et al., 1997;
Wu and Nancollas, 1998; Rosmaninho and Melo, 2006). Wu and Nancollas (1998) reported
that the effect of the interaction energies between titanium oxide coated surfaces (Anatase
and Rutile) and calcium phosphate deposits is due to the high value of the Lewis acid-base
and the electrostatic energies. This finding is somehow arguable, since i) the contribution of
the electrostatic energy in aqueous solutions can be considered as repulsive interaction energy
(Oliveira, 1997) or can be neglected (Visser, 2001), and ii) it was ambiguous how much the
contribution of the Lifshitz-van der Waals energy of these coatings to the total interaction
energies was. In addition, the observed fouling process of calcium phosphate may not have
been pure crystallization process. The presence of particulate material, e.g. insoluble calcium
phosphate may give rise to additional particulate fouling. Rosmaninho et al. (2007) showed
that increasing γ 2− causes the surface reaction rate coefficient to decrease as shown in Figure
4.41, which in turn may decrease the deposition rate. Despite these results, Rosmaninho et al.
(2007) found that the total deposition rate is increased with increasing γ 2− . They concluded
that the adhesion coefficient dominates the deposition process rather than the surface reaction
rate as shown in Figure 4.41.

84
4 Experimental Results and Discussion

surface reaction rate


1.6 adhesion coefficient 1.6

kr [m .kg/s] 1.2 1.2


TiN4

ka [m/s]
4

TiN2
0.8 0.8

TiN5 TiN3 TiN1


0.4 0.4

0.0 0.0
15 20 25 30 35 40 45 50 55 60
− 2
γ2 [ mJ/m ]

Fig.4.41.Surface reaction rate and adhesion coefficients as a function of electron donor


component for TiN modified surfaces (Rosmaninho et al., 2007)

However, this correlation does not apply to some coated surfaces such as SB2, WB,
Ni-P and K-S2. Even though the WB and Ni-P coated surfaces have higher values of γ 2−
which are 14.65 and 16.64 mJ/m2, nevertheless their initial fouling rates were found to be
higher. The weak abrasion resistance of the WB coating, which is only about 10% of that of
the SB coatings, may account for this behaviour. Furthermore, it was observed that parts of
the coating layer flaked off the surface, once the crystals started to deposit. As a consequence,
the damaged surface provides additional nucleation sites which result in increased fouling
rates. Figure 4.42 presents a picture of the WB coated surface which was taken after
terminating the experiment.

WB

flaking-off areas

Fig. 4.42 WB coated surface after fouling test

85
4 Experimental Results and Discussion

The surface was cleaned with a dilute solution of HCl and then rinsed with deionized
water. Though not clearly visible due to the weak colour contrast of the coating, some spots
are indicated by arrows which represent places where the flaking-off has taken place.
Figure 4.43 illustrates how the induction time varies with γ 2− . It is apparent that i)

only a marginal variation in induction time occurs if the γ 2− changed only slightly, e.g. for SB

coatings, and ii) for a significant change of the γ 2− , longer induction times have been
observed.

600
Ni-P-BN

500

400
tind [min.]

300

200

SB1 K-S1
100
SB4
SB3
0 AISI 304 BA

0 5 10 15 20 25
- 2
γ2 [mJ/m ]

Fig. 4.43 Variation of induction time with the electron donor component for untreated and
modified surfaces

4.2.5 Reproducibility of fouling propensity


Reproducibility of the fouling experimental data used to ascertain the fouling
behaviour of heat transfer surfaces under the same operating conditions. It can be established
by comparison of fouling rate and induction time for two identical trials with the same
operating conditions. To do so, two identical surfaces were investigated in two trials at the
same operating conditions. The experimental results as shown in Figures 4.44 and 4.45
indicate that:

86
4 Experimental Results and Discussion

i) the SB coated surfaces showed approximately similar crystallization fouling


behaviour except for SB2 coated surface. In other words, the fouling propensity of
SB2 coated surface of the second trial is inconsistent with those of the first trial,
ii) in case of WB coated surface, a relatively higher fouling rate and undoubtedly
shorter induction time were the main features of the second trial. This can be
attributed to the instability of the coating layer (low abrasion resistance),
iii) the reliable performance of the Ni-P coated surface was unsatisfactory, since
higher fouling rate and shorter induction time occurred for the second trial
compared with the values of the first trial. No plausible interpretation for such
behaviour has been found.

8
st
1 trial
initial fouling rate [10 m K/W. min]

7 nd
2 trial

6
2

5
-6

2 o
Tb=40 C
v = 0.15 m/s
1
cb= 4 g/L

0
SB1 SB2 SB3 SB4 WB Ni-P
surface

Fig. 4.44 Reproducibility in terms of initial fouling rate


for the modified surfaces

200 st o
1 trial Tb=40 C
nd
2 trial v = 0.15 m/s
180
cb= 4 g/L
160
induction time, min

140

120

100

80

60

40

SB1 SB2 SB3 SB4 WB Ni-P


surface

Fig. 4.45 Reproducibility in terms of induction time for


the modified surfaces

87
5 Theoretical Study

Chapter 5
THEORETICAL STUDY

The theoretical part of the present study endeavours to develop a new fouling criterion
that predicts whether a surface will foul or not. To do so, firstly a comparison will be made
between the previous criteria/models and present experimental results. Afterwards the
contribution of individual components of total surface energy will be investigated followed
by the development of a new theoretical criterion. Finally the chapter discusses the
comparison between predicted and experimental results as well as to those of previous
investigations.

5.1 Validation of the Previous Surface Energy Fouling Models


Förster (2001) developed a fouling criterion for predicting the induction time in terms
of surface texture and surface energy effects. He stated that a minimum value of spreading
coefficient ( λ12 ) and a maximum surface texture (Kf) are required to achieve minimum
fouling occurrence. On the other hand, a high value of λ12 corresponds to a high adhesion
strength and consequently a short induction time. Table 5.1 compares experimental and
predicted values of induction time, apparently a high degree of error exists.

Table 5.1 Comparison of predicted and experimental induction time of investigated


surfaces based on Förster (2001).
Surface Λ Kf/1010 m-2 Induction time [h]
predicted values experimental
based on Eq. (2.20) results
AISI 304 BA 1.04 0.087 9.4 0.72
SB1 1.54 0.080 45 1.48
SB3 1.42 0.065 37 0.97
SB4 1.45 0.055 39 1.20
K-S1 1.78 0.614 190 2.16
Ni-P-BN 0.72 0.45 22 9.66

The comparison of experimental results in terms of initial fouling rate and the
interfacial energy ΔG f which is calculated by Eq. (2.21) also yields no clear relation between

88
5 Theoretical Study

these two parameters, since high fouling rates occurred for surfaces having high value of
interfacial energy. The discrepancy of comparison may be related to omission of the polar
component to the interfacial energy.

-6
5.0x10
initial fouling rate [m K / W min]

AISI 304 BA

-6 SB1
4.0x10 SB4
SB3
2

-6
3.0x10
K-S1

-6
2.0x10

-6 Ni-P-BN
1.0x10

-108 -106 -104 -102 -100 -98


2
interfacial energy (ΔGf) [mJ/m ]

Fig. 5.1 Initial fouling rate as a function of interfacial energy for modified and untreated
surfaces based on Kwang-Ho and Chung-Hak (2000)

Zhao and Müller-Steinhagen (2002) found that, for CaSO4 deposition, a modified
surface having γ 2LW equal to 28 mJ/m2 will not foul or will have an infinite induction period.
Such phenomenon seems to be not valid for the investigated surfaces in this work. Table 5.2
underlines low accuracy between the experimental and predicted values in terms of induction
time

Table 5.2 Comparison of predicted and experimental induction time of investigated


surfaces based on Zhao and Müller-Steinhagen (2002)
Surface γLW, tind [h], predicted values tind, experimental,
[mJ/m2] based on Eq. (2.26) [h]
AISI 304 BA 35.11 160 0.72
SB1 23.73 234 1.48
SB3 27.36 1443 0.97
SB4 24.3 270 1.20
K-S1 21.36 148 2.16
Ni-P-BN 45.55 69 9.66

89
5 Theoretical Study

The comparison also underlines that i) surfaces having γ 2LW approximately equal to
28 mJ/m2 such as SB3 showed a finite period of induction time, and ii) there is no significant
reduction in induction time with increasing γ 2LW as would have been expected from the
criteria.

5.2 A New Simplified Surface Energy Fouling Criterion


In section 4.2 the behaviour of various modified surfaces when subjected to fouling
was analyzed based on i) the variation of fouling resistance with time and ii) the morphology
of the deposit layer. In addition, the experimental results in terms of initial fouling rate and
induction time have been interpreted according to the impact of surface energy components.
It has been found that γ 2− systematically influences the effect of modified surfaces on fouling
mechanisms.
Based on the DLVO theory the adhesion process is closely linked to the
intermolecular interaction energies between a deposit and a surface. To develop a new
criterion for predicting the tendency of a surface to fouling the following assumptions are
made:
i) it has previously been established that the contribution of the electrostatic energy in
aqueous solutions, particularly water, can be considered as repulsive interaction energy, since
all surfaces tend to acquire a like charge (Oliveira, 1997). In addition, the constant values of
bulk concentration and pH would definitely lead to the stability of the electrostatic double
layer energy, since it is strongly sensitive to the variation of these values (Israelachvili,
1991). However, Zhao and Müller-Steinhagen (2002) calculated the electrostatic double layer
energy for stainless steel and different modified surfaces which were subjected to the fouling
from calcium sulphate solutions. They found that for a given particle radius (R) and
separation distance (H) the electrostatic energy has a very small and constant value for all
surfaces. It was about 0.478 x 10-9 mJ/m at H equal to 0.157 nm. It is therefore plausible to
ignore the contribution of this energy to the total interaction energy,
ii) the Brownian motion is independent of surface energy and it has a small and
constant value as the bulk temperature is kept unchanged for all experiments. It was about
0.431 x 10-17 mJ. In such a case, the Brownian motion can also be neglected. Therefore the
total interaction energy, ΔE132
TOT
, will consist of only the other two energies, i.e. the Lifshitz-
van der Waals and the Lewis acid-base energy, thus Eq. (2.14) yields:

90
5 Theoretical Study

ΔE132
TOT
= ΔE132
LW
+ ΔE132
AB
(5.1)

The Lewis acid-base energy ΔE132


AB
as given in Eq. (2.12) is mainly a function of i)

ΔE132
AB
( H o ) which represents the free energy at the equilibrium distance (Ho), and ii) the actual

distance (H) between interacting bodies. Basically, Ho is defined as the minimum equilibrium
distance between the two interacting bodies, which has been found for a large range of
materials to be equal to 0.157 nm (van Oss, 1994).
iii) In case of H to be equal to Ho, ΔE132
AB
would only be a function of ΔE132
AB
( H o ) and

thus, of γ 2+ and γ 2− components of the interacting bodies. If a deposit (1) interacts with a
heated surface (2) when immersed in a polar medium, such as water (3) as shown in Figure
5.2, then ΔE132
AB
can be expressed as

⎛ γ +γ − + γ +γ − − γ +γ − + ⎞
⎜ 1 3 2 3 3 3 ⎟
⎜ − + ⎟
ΔE132
AB
= 4π R λ ⎜ γ 1 γ 3 + γ 2 γ 3 − γ 3 γ 3 − ⎟
− + + −
(5.2)
⎜ + − ⎟
⎜ γ 1 γ 2 − γ 1− γ 2+ ⎟
⎝ ⎠

γ3
liquid (3)

crystal (1) θ γ2
γ12
solid (2)

Fig. 5.2 Crystal, surface and liquid interfaces energies

As a result, Eq. (5.1) can be expressed as

⎛ γ +γ − + γ +γ − − γ +γ − + ⎞
⎜ 1 3 2 3 3 3 ⎟
ΔE
{( )}
TOT
⎜ − + ⎟ (5.3)
132
= H o 2 γ 1LW γ 3LW + γ 2LW γ 3LW − γ 1LW γ 2LW − γ 3LW + 2λ ⎜ γ 1 γ 3 + γ 2 γ 3 − γ 3 γ 3 − ⎟
− + + −

C
⎜ + − ⎟
⎜ γ 1 γ 2 − γ 1− γ 2+ ⎟
⎝ ⎠

91
5 Theoretical Study

To asses the assumption that H is equal to Ho, the data for a stainless steel surface
immersed in an aqueous suspension of calcium sulphate were used to evaluate the ΔE132
LW
and

ΔE132
AB
energies. Figure 5.3 demonstrates that ΔE132
AB
is approximately zero for H greater than

5 nm. This is plausible as ΔE132


AB
is a short-range energy operating at a distance less than 0.3

nm (Oliver, 1997). In addition, the difference between ΔE132


LW
and ΔE132
AB
energies seems to be
marginal with increasing separation distance between the attractive bodies. On the other
hand, the range between these two interfacial energies is too high at low separation distance,
i.e. at 0.157 nm. Based on these two observations, it is reasonable to study the influence of
these interaction energies at the region where H equals to Ho.

-9
1x10

0
0 2 4 6 8 10
ΔΕ132 /C [mJ/m]

separation distance (H) [nm]


-9
-1x10

LW
-9
Lifshitz-van der Waals energy, ΔE132
-2x10 AB
Lewis acid-base energy, ΔE132

-9
-3x10

Ho= 0.157 nm

Fig. 5.3 Lifshitz-van der Waals and Lewis acid-base interfacial energies as a function of
separation distance

The calculated results for individual terms in Eq. (5.3) are plotted in Figure 5.4.
Obviously, the contribution of ΔE132
AB
to the total interaction energy seems to be dominant. In

particular, the influence of ΔE132


LW
for the various coatings on the total interaction energy is

marginal, leaving ΔE132


AB
to be the most dominant term.

92
5 Theoretical Study

-9
4x10 LW
Lifshitz-van der Waals energy, Δ E132
AB Ni-P-BN
3x10
-9 Lewis acid-base energy, Δ E132
TOT
Total interaction energy, Δ E132
-9
2x10
K-S1
-9
1x10
ΔE132/C [mJ/m]

SS SB1 SB3 SB4


0

-9
-1x10
-9
-2x10
-9
-3x10
-9
-4x10
-9
-5x10

Fig. 5.4 Contribution of Lifshitz-van der Waals and Lewis acid-base energies to the total
interaction energy

Based on these results, Eq. (5.3) can be simplified to:

⎛ γ +γ − + γ +γ − − γ +γ − + ⎞
⎜ 1 3 2 3 3 3 ⎟
ΔE132
TOT
⎜ − + ⎟
= 2λ ⎜ γ 1 γ 3 + γ 2− γ 3+ − γ 3+ γ 3− − ⎟ (5.4)
C
⎜ + − ⎟
⎜ γ 1 γ 2 − γ 1− γ 2+ ⎟
⎝ ⎠

However, van Oss et al. (1987) stated that it suffices to use polarity ratios of γ i+ and

γ i− relative to the values of water. The polarity ratios can then be defined as follows;

δ i+ = γ i γ i−
+
and δ i− = . δ i+ and δ i− are the relative Lewis acid and base polarities
γ +
3 γ −
3

γ 3+γ 3−
of substance i with respect to water. Multiplying the right hand side of Eq. (5.4) by
γ 3+γ 3−
yields

93
5 Theoretical Study

⎛ γ 1+ γ 3− γ 2+ γ 3− γ 3+ γ 3− γ 1− γ 3+ ⎞
⎜ + − + +⎟
⎜ γ 3+ γ 3− γ 3+ γ 3− γ 3+ γ 3− γ 3+ γ 3− ⎟
ΔE132
TOT
+ −⎜ ⎟
= 2λ γ 3 γ 3 (5.5)
C ⎜ γ 2− γ 3+ γ 3+ γ 3− γ 1+ γ 2− γ 1− γ 2+ ⎟
⎜ − − − ⎟
⎜ γ 3+ γ 3− γ 3+ γ 3− γ 3+ γ 3− γ 3+ γ 3− ⎟
⎝ ⎠

Simplifying the
γ i+ and
γ i− terms to δ i+ and δ i− , respectively, gives rise to:
γ +
3 γ 3−

ΔE132
TOT

C
(
= 51λ δ 1+ + δ 2+ + δ 1− + δ 2− − δ 1+ δ 2− − δ 1−δ 2+ − 2 ) (5.6)

Having a closer look at Eq. (5.6), it can be clearly seen that the total interaction
energy depends only on the contribution of the δ i+ and δ i− components of the interacting
bodies. However, van Oss (1994) assumed that i) a solid surface or a foulant can be
considered as a monopolar base substrate when δ + < 0.2 and ii) when δ + of the solid surface

and of the foulant are significantly greater than 0.2, then the state is called a bipolar system.
Based on these assumptions, it is possible to express Eq. (5.6) as:

ΔE132
TOT
= a + bδ 2− (5.7)
C
Where

(
⎧ δ 1− − 2 ) monopolar system ( solid surface and foulant )
⎪ +
a = 51λ ⎨ +
(
⎪ δ1 + δ1 − 2

) monopolar solid surface
(
⎪ δ 2 + δ1 − δ 2 δ1 − 2
− + −
) monopolar foulant
⎩ 1 (
⎪ δ + + δ + + δ − − δ +δ − − 2
2 1 2 1 ) bipolar system ( solid surface and foulant )

⎧1 monopolar system or monopolar foulant


b = 51λ ⎨
⎩(1 − δ 1 )
+
bipolar system or monopolar surface

Equation (5.7) shows that increasing the Lewis base polarity δ i− would make the total
interaction energy to be more repulsive than attractive. On the other hand, the higher the

94
5 Theoretical Study

electron donor component the higher the energy barrier of adhesion would be. As a
consequence, lower fouling rates are expected for surfaces having higher electron donor
components.

5.3 Validation of the New Surface Energy Fouling Criterion


To validate the new surface energy fouling criterion for predicting the fouling
propensity of different surfaces, it is firstly compared with experimental results of the present
study. As mentioned in Chapter 4, the SB coated surfaces reduce the Lifshitz-van der Waals
component ( γ 2LW ) by 20-46% in comparison to the stainless steel surface. Such reduction

causes the Lifshitz-van der Waals energy ( ΔE132


LW
) to be less attractive than the original
stainless steel surface as shown in Figure 5.5. However, the Ni-P-BN coating increases the
ΔE132
LW
attractive energy while its fouling rate is even lower than that for the other surfaces

(SB and stainless steel). This indicates that ΔE132


LW
is inadequate to interpret the adhesion

process. This is in agreement with the results shown in Figure 5.4, i.e. the influence of ΔE132
LW

for the various coatings on the total interaction energy is marginal.

-10
2x10

K-S1
0
SB1
SB4
-10
-2x10
ΔE132 /C [mJ/m]

SB3
-10
-4x10
AISI 304 BA
LW

-10
-6x10

-10
-8x10 Ni-P-BN

-9
-1x10
15 20 25 30 35 40 45 50
LW 2
γ2 [mJ/m ]

Fig. 5.5 Lifshitz-van der Waals energy as a function of γ 2LW of the different investigated
surfaces

95
5 Theoretical Study

To characterise the deposition process based on surface properties, the other


interfacial energy "Lewis acid-base energy" seems to be more relevant. This is because all
coatings display a systematic increase in the electron donor component of the surface energy,
which in turn corresponds to increased Lewis acid-base energy. Under such circumstances,
the surfaces are rather more repulsive than attractive as shown in Figure 5.6. It can therefore
be concluded that an increase in the electron donor component (Lewis base polarity) causes
ΔE132
TOT
to increase, which in turn promotes the repulsive interaction energy between the
surface and the deposits. This means that lower fouling rates are expected for surfaces having
higher electron donor components. This is in fact consistent with the expected results from
Eq. (5.7).

-9
4.0x10
Ni-P-BN

-9
2.0x10
K-S1
ΔE132/C [mJ/m]

0.0
SB1
AB

SB3
-9
-2.0x10
SB4

-9
AISI 304 BA
-4.0x10

0 5 10 15 20 25 30
− 2
γ2 [mJ/m ]

Fig. 5.6 Lewis acid-base energy as a function of electron donor component of the investigated
surfaces

If in Eq. (5.7) δ 2+ of the modified surfaces and of the foulant (calcium sulphate, δ 1+ )
are less than 0.2 as shown in Table 5.3, the terms "a" and "b" can be considered as a
(
monopolar system, i.e. a = 51λ δ 1− − 2 and b = 51λ . )

96
5 Theoretical Study

Table 5.3 Lewis acid polarity of the untreated and modified surfaces
Surface Lewis acceptor Lewis acid polarity
component,
(mJ/m2)
AISI 304 BA 0.03 0.0343
SB1 0.01 0.0198
SB3 0.0021 0.0091
SB4 0.05 0.0442
K-S1 0.003 0.0108
Ni-P-BN 0.45 0.1328
CaSO4.2H2O (Förster, 2001) 0.02 0.0280

ΔE132
TOT
To minimize fouling where > 0, it is essential to create a situation where δ 2− >
C
0.685. This means that the deposition of calcium sulphate is not expected to occur for
surfaces having γ 2− greater than 12 mJ/m2. On the other hand, lower fouling rates are
expected for surfaces having higher electron donor components compared to that of AISI 304
BA as clearly demonstrated in Figure 5.7.

5
AISI 304 BA
initial fouling rate [10 m K/W.min]

SB1
4
SB4
SB3
K-S1
2

3
-6

1
Ni-P-BN

0
-4.0x10
-9 -9
-2.0x10 0.0 2.0x10
-9 -9
4.0x10
TOT
ΔE132 /C [mJ/m]

Fig. 5.7 Initial fouling rate as a function of the total interaction energy

97
5 Theoretical Study

The new criterion has also been validated with the data of the previous studies of
Förster (2001) and Zhao et al. (2009) for various foulants (CaSO4 and bacteria). The selection
of these studies was due to the available information on both the electron donor and electron
acceptor components, while many previous studies did not report these components, since
different approaches rather than the Lewis acid-base approach have been used for
determining the surface energy components. The data presented by Rosmaninho (2007) were
not selected for comparison with the new criterion as the experimental data was a result of
mixed crystallization and particulate fouling.
Förster (2001) studied the effect of surface energy of polymer coated surfaces and
metallic surfaces on fouling mechanisms of calcium sulphate under convective heat transfer
conditions. The surface energy components as well as the experimental results in terms of
initial fouling rate are illustrated in Appendix A3. From these data one can observe that all of
the modified surfaces have δ 2+ < 0.2. In addition, the foulant (CaSO4.2H2O) has also δ 1+ < 0.2.
Thus, the terms "a" and "b" are considered as a monopolar system. To minimize the adhesion
between the deposits and the surfaces (repulsive interaction energy), it is therefore essential
to provide surfaces with γ 2− greater than 12 mJ/m2. As an expected, the lower initial fouling
rate occurred on the bronze surface where repulsive interaction energy has been achieved as
shown in Figure 5.8. It is noteworthy that the polymer coated surfaces (PTFE and PFA) and
the metallic surfaces (copper and brass) do not follow the expected trend. This can be
attributed to the following reasons:

i) the copper surface has δ 2+ > 0.2. Therefore, the terms "a" and "b" in Eq. (5.7)

should be considered as a monopolar foulant. In such a case surfaces with γ 2−


greater than 18 mJ/m2 are necessary to reduce the adhesion energy. In other
words, surfaces having γ 2− less than 18 mJ/m2 such as copper are more susceptible
to fouling due to the attractive interaction energy between the deposit and the
surface.
ii) the brass, PTFE and PFA surfaces have a zero electron donor component which
in turn leads to eliminate the contribution of the Lewis acid-base energy to the
total interaction energy.

98
5 Theoretical Study

2.4
o
Tb = 40 C
Cu
initial fouling rate [10 m K/W h] v = 0.15 m/s
2.0 cb = 4 g/l
2

Ms(I)
1.6
-5

1.2
PTFE

FEP
0.8 PFA Al

Ms(II) Bronze
0.4

-8.0x10
-9
-6.0x10
-9
-4.0x10
-9
-2.0x10
-9
0.0 2.0x10
-9

TOT
ΔE132 /C [mJ/m]

Fig. 5.8 Initial fouling rate based on Förster (2001) versus total interaction energy of metallic
and polymer coated surfaces

The impact of surface energy on biofouling has been investigated by Zhao et al.
(2009). Si and N-doped DLC films with various Si and N contents were deposited on glass
slides using magnetron sputter ion-plating and plasma-enhanced chemical vapour deposition
(PECVD). The surface energy components as well as the experimental results in terms of
adhered bacteria are illustrated in Appendix A4. It can be seen that the DLC1, DLC4, DLC5
and DLC6 modified surfaces are bipolar substrates, since δ 2+ > 0.2 ,while in case of DLC2,

DLC3 and DLC7 modified surfaces the δ 2+ is less than 0.2 which causes these surfaces to
have monopolar characterizations. In addition, the foulant (Pseudomonas Fluorescens) has a
δ 1+ > 0.2. Thus the terms "a" and "b" become

⎧⎪ δ + + δ 1− − 2
a = 51λ ⎨ 1+
( )
DLC 2, DLC 3 and DLC 7 " monopolar solid surface"
( )
⎪⎩ δ 2 + δ 1+ + δ 1− − δ 2+ δ 1− − 2 DLC1, DLC 4, DLC 5 and DLC 6 " bipolar system"

b = 51λ (1 − δ 1+ )

99
5 Theoretical Study

To minimize fouling the following conditions should be met:

γ 2− > 1.85 mJ/m2 for DLC2, DLC3 and DLC7

γ −
>
(0.66 δ +
− 0.138
2 )
mJ/m2 for DLC1, DLC4, DLC5 and DLC6
2
0.516

For DLC1, DLC4, DLC5 and DLC6 the essential values of γ 2− to minimize fouling

vary with respect to the value of γ 2+ of the modified surfaces. Table 5.4 shows the theoretical

minimum values of γ 2− required to increase the energy barrier of the adhesion process.

Table 5.4 Theoretical minimum values of γ 2−


for DLC1, DLC4, DLC5 and DLC6 modified surfaces
Coating γ 2− [mJ/m2]
DLC1 2.46
DLC4 0.25
DLC5 0.17
DLC6 0.0034

Figure 5.9 demonstrates that DLC4, DLC5 and DLC6 modified surfaces reduce the
adhering bacteria significantly in comparison to the DLC1 modified surface. These results
support the hypothesis that reduced bacterial adhesion is expected on surfaces having γ 2−

greater than the theoretical minimum value. For the DLC1 surface, γ 2− is less than the
theoretical value by approximately 36%, which in turn increases bacterial adhesion on the
surface. On the other hand, an increase in the electron donor component causes ΔE132
TOT
to
increase, which in turn promotes the repulsive interaction energy between the surface and the
bacteria.
For the DLC2 and DLC3 modified surfaces, in spite of having higher electron donor
component than the theoretical value (1.82 mJ/m2), the adhering bacteria count was found to
be higher than expected value (see Figure 5.9). No plausible interpretation for such behaviour
has yet been found. It may be speculated that the chemistry of these coatings may have an
influence on the adhesion process of Pseudomonas Fluorescens.

100
5 Theoretical Study

5000
monopolar surfaces
DLC1
Adhered bacteria [CFU/cm ]
DLC3
2

4000
DLC2

3000

2000
DLC5 DLC7

DLC4
1000
bipolar system DLC6

0 0 -9 -9 -9 -9
2x10 4x10 6x10 8x10

TOT
ΔE132 /C [mJ/m]

Fig. 5.9 Amount of adhered bacteria according to Zhao et al. (2009) as a function of the total
interaction energy of the modified surfaces

These comparisons confirm that this new criterion can potentially be used as a
predictive tool to see if surface would foul or not. The tendency of surfaces to foul decreases
with increasing γ 2− (Lewis base polarity). In other words, a higher electron donor component
is indicative of weaker attractive interaction between the surface and the deposits. Thus, to
minimize fouling it is essential to create surfaces having higher γ 2− .

101
6 Field Investigation of Coated Surfaces

Chapter 6
FIELD INVESTIGATION OF COATED SURFACES

As stated in Chapter 1, the investigation of innovative anti-fouling coatings was the


main contribution of ITW to the MEDESOL project. A maintenance free heat exchanger is
sought as the ultimate desalination plant to be installed in rural areas. Accordingly, it was
appropriate to investigate the coated surfaces with promising laboratory results in the
desalination plant of MEDESOL. This chapter details specifications of the plate heat
exchanger as well as field analysis of the coated plates when subjected to saline water.

6.1 Heat Exchanger Specifications


A schematic diagram of the plate heat exchanger is presented in Figure 6.1. The
bolted heat exchanger contained 10 thin corrugated plates. The heating fluid (distilled water)
comes from the solar collectors while the heated fluid is saline water. The design parameters
of the heat exchanger are given in Table 6.1.

Fig. 6.1 Schematic diagram of the plate heat exchanger

102
6 Field Investigation of Coated Surfaces

Table 6.1 Heat exchanger design parameters


Properties Hot side Cold side
maximum inlet temperature (oC) 95 55
nominal outlet temperature (oC) 80 85
pressure drop (kPa) 49.38 21.32
mass flow rate (kg/s) 1.11 0.58
heat exchange rate (kW) 70 70

6.2 Examination of the Modified Surfaces in the MEDESOL-I Prototype


The main conclusion of the preceding chapters was that surfaces with high γ 2−
produce a higher reduction in the total interfacial energy which causes the adhesion energy
between the surface and the deposits to decrease. Higher values of γ 2− also indicate easier
removal of the deposits from the heat transfer surface. Accordingly, SB1 and SB3 coatings
that performed best in terms of overall fouling mitigation were chosen for coating and
examining the plates. The maximum allowable temperature of these coated surfaces is 200oC
which is far above the highest operating temperature in the MEDESOL project (95oC).
Moreover, in order to maximize the use of the MEDESOL infrastructures and time, it has
been decided to evaluate the performance of SB1 and SB3 coated plates simultaneously.
Accordingly, plates 1-5 were coated with SB1 and plates 6-9 with SB3. Plate no. 10 remained
uncoated. Furthermore, gasket areas of the plates were not coated due to hydrophobicity of
the coating layer which otherwise would have resulted in leakage (see Figure 6.2). Typical
coated areas and uncoated gasket areas are shown in Figure 6.3.

Fig. 6.2 Uncoated plates with covered gasket area

103
6 Field Investigation of Coated Surfaces

Fig. 6.3 Coating of a typical plate with gasket area remaining uncoated

The objective of the field investigation was not only to assess the thermal
performance but also the mechanical stability of the coatings. The plate heat exchanger was
subjected initially to synthesized calcium sulphate solution (as defined in MEDESOL
prototype-I). The typical temperature difference between the hot and cold sides was in the
range of 40-45oC.
The experiments were done at the Plataforma Solar de Almeria in Almeria, Spain
(PSA), and the following conditions were adjusted:

ƒ The heat exchanger was in operation in various time intervals of 8 hours for
almost 45 days with a CaSO4 concentration of 4 g/L which replicates laboratory
test solution at ITW.
ƒ The flow rates were maintained almost constant at 20 l/min
ƒ The heating temperature varied from 20 to 85ºC in accordance with the solar
irradiation

After 45 days of experiments, the heat exchanger was dismantled. Afterwards the
plates were examined to detect any sign of fouling, as shown in Figure 6.4. Apparently, there
is no sign of fouling, especially for SB1 coated plates. These findings confirm and support
the laboratory results obtained at ITW, see Figure 4.27 and Table 4.4. Although no major
sign of scaling was found, some of the coatings had peeled off, especially near the inlet and

104
6 Field Investigation of Coated Surfaces

outlet ports (see Figure 6.5). One side of the stainless steel plate No. 10 was only subjected to
the heating fluid (distilled water), while the other side was in the dead zone where there was
not any flow.

Fig. 6.4 Performance of plate 4 (SB1), cold side

Fig. 6.5 Performance of plate 5 (SB1), inlet areas

At the hot side where only distilled water was supposed to flow, no fouling should
have occurred as shown in Figure 6.6. Nevertheless a lot of damage had occurred to the
coating layer, particularly for the SB3 coated plates. Furthermore, one could easily remove
the remaining coating layer already by scrubbing it by hand. Such observation would indicate

105
6 Field Investigation of Coated Surfaces

the impact of thermal/mechanical stresses on the coatings, which were taken into account for
the later phase of the project.

Fig. 6.6 Performance of plate 7 (SB3), hot side

These results were discussed with the company responsible for the coating (ItN
Nanovation), and it was agreed to apply thinner layers of the same coatings to increase the
mechanical stability. Therefore, the plates were sent back to Germany where they were
recoated. Initially, the surfaces were sand blasted to increase the surface roughness as
common practice to realize stronger adhesive joints between the substrate and the coating
(Reegen and Ilkka, 1962; Packham, 2005). Afterwards, the surfaces were recoated with a
thinner layer of the same coatings.
In the second field trial, the plate heat exchanger was subjected to well water which
was slightly salty. The heat exchanger was in operation almost 30 days in time intervals of 8
hours per day. The hot side temperature varied from 20-90oC due to solar irradiation, and the
flow rate on the cold side was 20 L/min. From the post-fouling examination of the plates the
following conclusions were drawn:

ƒ Similar to the first trial, no sign of fouling was observed as shown in Figures 6.7
and 6.8.
ƒ A red discolourization has been observed which may have resulted from corrosion
of other components (storage tanks, pipes, etc.) as depicted in Figure 6.9.

106
6 Field Investigation of Coated Surfaces

Fig. 6.7 Performance of plate 1 (SB1), cold side

Fig. 6.8 Performance of plate 6 (SB3), cold side

Fig. 6.9 Discolourization of coating due to corrosion of other components of the test rig

107
6 Field Investigation of Coated Surfaces

ƒ The cold side where the saline water is heated was certainly the core interest of
the present investigation. Apart from the inlet and outlet areas and the contact
points between plates only very little peeling off has been observed for SB1
coated plates (Plates 1-5), see Figure 6.10. It has been stated that around the
contact points between the plates, the flow velocity is low, resulting in higher wall
temperatures (Bansal and Müller-Steinhagen, 1993). On the other hand, the
temperature profile in the vicinity of the contact points is different from that in the
flow channel, since a volume of liquid is heated from several directions. As a
result of these effects, most nucleation and crystal formation, if any, is initiated
near the contact points.

Fig. 6.10 Performance of plate 2 (SB1), cold side contact points

Through the results of the laboratory and the MEDESOL-I prototype one can
conclude that the adhesion energy between the deposits and the SB1 modified surfaces is so
weak that it can be overcome even by small shear forces. In addition, a small change in the
coating procedure i.e. a thinner layer of coating and slight roughening of the substrate can
result in substantial improvement in the mechanical stability of the coated plates.

108
7 Conclusions and Future Work

Chapter 7
CONCLUSIONS AND FUTURE WORK

7.1 Conclusions
Surface treatment has recently become the focus of much attention due to its
environmentally friendly features that limit the use of harmful chemicals to mitigate fouling. In
this study, fouling of calcium sulphate has been systematically investigated on various
modified surfaces during convective heat transfer. The operating conditions were similar to
those expected in a typical membrane distillation (MD) desalination plant. The experimental
part identified the influence of surface properties in terms of surface i) roughness, ii) texture,
and iii) energy related properties on crystallization fouling. In the theoretical part, a new
criterion is introduced that may predict under what circumstances a modified surface would
foul. The main conclusions that can be drawn from the experimental and theoretical results of
this study are:

7.1.1 Alteration of the surface geometry

Roughened surfaces
Sand papers with different grain sizes were used to provide roughenesses from 0.54 to
1.55 μm. The experimental results showed that, within the range of operating conditions, the
highest fouling rate as well as the shortest induction period were observed for the roughest
surfaces of 1.55 μm. Several facts contribute to this behaviour notably 1) the sharp edge of a
roughest surface could pin the triple line position far from a stable equilibrium state and
consequently reduce the contact angle. Decreasing the contact angle causes the nucleation
correction factor ( φ ) to decrease which possibly leads to decrease the energy barrier for
heterogeneous nucleation and consequently a higher fouling rate and 2) the rough surface
provides more cavities or pits which would act as nucleation sites so more precipitation is
expected.
In addition, for a given roughness value, increasing the supersaturation ratio causes a
significant increasing in fouling rate. This can be related to i) the increase in the driving
potential (Cb-Cs) and consequently enhanced nucleation rate (J) and ii) the reduction of the
double layer thickness of the electrostatic energy which in turn leads to a reduced energy

109
7 Conclusions and Future Work

barrier for adhesion, i.e. enhancement of adhesion between the deposits and the heat transfer
surfaces. The visual examination of the fouled surfaces indicated that the deposit layer on a
rough surface has a stronger adhesion than on a smooth surface and is generally thicker. The
results of this part were used in latter stages of the MEDESOL project as the coating
company used roughened surfaces to increase the stickiness of the coating to the substrate.

Structured surfaces
The advantageous aspects of enhanced or extended heat transfer surfaces may be
offset if subjected to fouling. In this study, plain stainless steel plates with 'V' shaped grooves
were used as heat transfer specimens and investigated under clean and fouling conditions. In
addition, the impact of i) direction of grooves with respect to fluid flow (crossed, longitudinal
and mixed flow grooves), and ii) the groove dimensions on fouling mechanisms has been
investigated. The experimental results highlighted that:

• For identical groove dimensions, clean heat transfer coefficients increased with
velocity for all directions, but more notably for the crossed ones.
• As for the crossed grooves, clean heat transfer is enhanced as both the depth and
width are increased. No significant impact of dimensions was observed for
longitudinal and mixed grooves.
• Of three different surface textures, the least deposition occurred for the crossed
grooves. The generated eddies within the crossed grooves may assist in removal
of initial crystals forming on the surface. The impact of these eddies is also
evident inside the crossed grooves, in which either no or only little deposit was
observed.

7.1.2 Alteration of the surface energy related properties


The alteration of surface energy related properties was used as an approach to
minimize foulant adhesion energies and provide an environmentally friendly non-stick
modified surface for utilization in the plate heat exchanger of the MEDESOL project. The
stainless steel surfaces were modified using different coating materials, namely solvent based
(SB) coatings, water based (WB) coatings, electroless Ni-P coatings and nano-structured
coatings. The surface energy of the modified surfaces was calculated based on the Lewis
acid-base approach. In comparison with stainless steel surfaces, the coating layers with

110
7 Conclusions and Future Work

thickness in the range of few micrometers significantly reduced the total surface energy by up
to 40%. Nevertheless, the most notable parameter among various surface energy components
was the electron donor component ( γ 2− ) that could systematically influence fouling

propensity i.e. higher γ 2− corresponded to lower deposition.


Fouling rate and induction time were used to quantify the resistance of modified
surfaces to fouling as well as to evaluate the impact of surface energy components on the
CaSO4 deposition process. The experimental results showed that for most modified surfaces
as time goes on the fouling resistance increases followed by an almost drastic drop implying
that the deposit layer has been spalled off the surface. This indicates that the low stickability
of the deposit layer to the surface together with increased shear forces of fluid can help
deposit removal. It can further be speculated that the remaining lumps of deposits will
increase the level of turbulence close to the heat transfer surface. This resulted in lower
deposition rates should crystals form on the surface again. It is noteworthy that for a modified
surface the impact of surface energy on fouling mechanisms overcomes the influence of
surface roughness. Nevertheless, no correlation between fouling rates of different modified
surfaces and the surface roughness was established, as has been done for bare stainless steal
surfaces.
Although tangible results were obtained for modified surfaces, their utilization in
industrial application is still problematic due to the relatively weak thermal, mechanical and
physicochemical stabilities of coatings. It has been found that the WB coating was unstable to
abrasion, since its coating layer flaked off the surface during the fouling run. For the Ni-P-
BN and the K-S2 coated surfaces some spots of oxidization were observed. The spots were in
red colour which is a feature of corrosion and occurred only at the interface between the
damaged part of the coating and the stainless steel base plate. It is not immediately clear how
and why this did happen. Nevertheless, it is worth to mention that no such occurrence was
observed when fouling tests were carried out on the uncoated stainless steel plate under
similar operating conditions. Among the investigated modified surfaces, the SB coatings,
with exception of SB2, showed better thermal/mechanical/physiochemical properties as well
as a reliable performance in terms of reproducibility.
In addition, deposit morphology analyses by SEM showed that deposit structure on
the different surfaces was also different in terms of size of crystals, orientation, hardness and
adherence. Small size and no defined orientation are the main characterizations of surfaces

111
7 Conclusions and Future Work

having high γ 2− . In addition, the deposits adhered more loosely and could easily be washed
away by just flushing the surface with diluted HCl solution, particularly the SB coated
surfaces.
As for the impact of surface energy components on the deposition process of calcium
sulphate, it was also found that:

• The total surface energy ( γ 2 ) as well as the Lifshitz-van der Waals component

( γ 2LW ) are not sufficient to describe the deposition process, since no


straightforward correlation between these two components and fouling propensity
in terms of induction and fouling rate has been found.
• The electron donor component seems to be a key factor that could be used in the
interpretation of the impact of the interfacial energies on the adhesion process of
the calcium sulphate deposits.

To quantify the adhesion process between the deposits and the modified surfaces,
some theoretical considerations have also been carried out to evaluate the contribution of
each individual intermolecular interaction energy as well as to develop a new criterion that
may predict the fouling tendency of modified surfaces. It was found that the contribution of
the Lewis acid-base energy to the total interaction energy between calcium sulphate deposits
and the modified surfaces exceeds the contribution of the Lifshitz-van der Waals energy.
Surfaces having higher electron donor component (Lewis base polarity) provide a higher
repulsive energy which causes the adhesion force between the surface and the deposits to
decrease, i.e. the electron donor component is a key parameter that may better correlate
fouling tendency and surface properties than surface energy itself. In addition, higher values
of γ 2− indicate easier removal of the fouling layer from the heat transfer surface.
Based on these findings, the solvent based coatings of SB1 and SB3 were chosen for
the modification of surface energy properties of the corrugated plate heat exchanger of the
MEDESOL project. Once coated, the heat exchanger was subjected to synthesized calcium
sulphate solution with a solution concentration of about 4 g/L and heating temperatures
ranging from 20 to 85ºC. The field results showed no sign of fouling thus validating the
suitability of the selected coatings for this application. However, the main drawback was the
weak adhesion of the coatings to the substrate. To overcome this in another attempt, the

112
7 Conclusions and Future Work

surfaces were slightly roughened before recoating with the same coatings but with reduced
thickness. In the second trial, the plate heat exchanger was subjected to saline well water for
30 days in time intervals of 8 hr/day. The heating temperature varied from 20 to 90oC. The
results of this trial were satisfactory as in addition to the avoidance of fouling, the stickiness
of the coatings to the substrate was much improved even at the contact points between the
plates as well as at the inlet and outlet ports.

7.2 Future Work


The experimental and theoretical results of this study highlighted the potential
advantages of modified surfaces to mitigate fouling, as well as challenges that require further
investigation. Some general remarks and suggestions for future work are as follows:

• In addition to the higher driving potential force of crystallization with increased


concentration, the electrostatic double layer energy will decrease due to a decrease
1
in Debye length ( ). In this study, the electrostatic energy was considered
κ
constant since all the experiments were carried out under constant foulant
concentration. Therefore, further experiments are required to discern the extent of
reduction in the electrostatic energy when foulant concentration increases.
• There is little information available on the role of the electrostatic double layer
energy in the fouling process. Such investigation would help to understand the
contribution of this energy to the total interaction energy.
• In case of grooved surfaces and in some cases of modified surfaces the fouling
curves level off after a certain period of time. Bubble generation within the pores
of the deposit layer due to high temperature at the deposit/solid interface is
considered to be a possible reason behind this phenomenon. Further investigations
especially with the help of a high-speed camera would certainly shed some light
on this peculiar trend.

113
List of References

LIST OF REFERENCES

Augustin, W., Geddert, T., Scholl, S., 2007, Surface treatment for the mitigation of whey
protein fouling, Proceeding of Heat Exchanger Fouling and Cleaning: Challenges and
Opportunities, Tomar, Portugal, Vol. RP5, pp. 206-214.

Augustin, W., Zhang, J., Bialuch, I., Geddert, T., Scholl, S., 2005, Modified DLC-coatings
for the mitigation of scaling on heat transfer surfaces, Proceeding of International
Conference on Enhanced, Compact and Ultra-Compact Heat Exchangers: Science,
Engineering and Technology, Whistler, British Columbia, Canada, pp. 394-402.

Bansal, B., Müller-Steinhagen, H., 1993, Crystallization fouling in plate heat exchangers,
Journal of Heat Transfer, Vol. 115, pp. 584-591.

Bansal, B., Müller-Steinhagen, H., Dong, X., 2000, Performance of plate heat exchangers
during calcium sulphate fouling investigation with an in-line filter, Chemical Engineering
and Processing, Vol. 39, pp. 507-519.

Benzinger, W., Schygulla, U., Jäger, M., Schubert, K., 2005, Anti fouling investigations with
ultrasound in a micro-structured heat exchanger. Proceeding of Heat Exchanger Fouling and
Cleaning: Challenges and Opportunities, Kloster Irsee, Germany, Vol. RP2, pp. 197-201.

Beuf, M., Rizzo, G., Leuliet, J. C., Müller-Steinhagen, H., Yiantsios, S., Karabelas, A.,
Benezech, T., 2003, Fouling and cleaning of modified stainless steel plate heat exchangers
processing milk products, Proceeding of Heat Exchanger Fouling and Cleaning:
Fundamentals and Applications, Santa Fe, New Mexico, USA, Vol. RP1, pp. 99-106.

Bohnet, M., 1987, Fouling of heat transfer surfaces, Chemical Engineering Technology, Vol.
10, pp. 113–125.

Bornhorst, A., Müller-Steinhagen, H., Zhao, Q., 1999, Reduction of scale formation under
pool boiling conditions by ion implantation and magnetron sputtering on heat transfer
surfaces, Heat Transfer Engineering, Vol. 20, pp. 6-14.

114
List of References

Bott, T. R., 1995, Fouling of Heat Exchanger, Elsevier Science B. V., The Netherlands.
Carlson, J. A., 1992, Understanding the capabilities of plate-and-frame heat exchangers,
Chemical Engineering Progress, Vol., 59, pp. 26-31.

Chen, H., Chen, D., Li, Y., 2006, Investigation on effect of surface roughness pattern to drag
force reduction using rotary rheometer, Journal of Tribology, Vol.128, pp. 131-138.

Cho, Y. I., Lee, S., Kim, W., Suh, S., 2003, Physical water treatment for the mitigation of
mineral fouling in cooling-tower water applications, Proceeding of Heat Exchanger Fouling
and Cleaning: Fundamentals and Applications, Santa Fe, New Mexico, USA, Vol. RP1, pp.
20-31.

Derjaguin, B. V., Landau, L., 1941, Acta physicochem, URSS, Vol. 14, pp. 633-662.

Dukhin, S. S., Derjaguin, B. V., 1974, Electrokinetic Phenomena, J. Willey and Sons.

Eustathopoulos, N., Nicholas, M. G., Drevet, B., 1999, Wettability at High Temperatures,
Pergamon, Elsevier Science Ltd., Kidlington, Oxford, UK.

Förster, M., 2001, Verminderung des Kristallisationsfoulings durch gezielte Beeinflussung


der Grenzfläche zwischen Kristallen und Wärmeübertragungsfläche, PhD dissertation,
Fakultät für Maschinenbau und Elektrotechnik, Universität Carolo-Wilhelmina zu
Braunschweig, Germany.

Förster, M., Augustin, W., Bohnet, M., 1999, Influence of the adhesion force crystal/heat
exchanger surface on fouling mitigation, Chemical Engineering and Processing, Vol. 38, pp.
449-461.

Förster, M., Bohnet, M., 1999, Influence of the interfacial free energy crystal/heat transfer
surface on the induction period during fouling, International Journal of Thermal Science,
Vol. 38, pp. 944-954.

115
List of References

Förster, M., Bohnet, M., 2000, Modification of molecular interactions at the interface
crystal/heat transfer surface to minimize heat exchanger fouling, International Journal of
Thermal Science, Vol. 39, pp. 697-708.

Förster, M., Bohnet, M., 2001, Modification of interface crystal/heat transfer surface to
reduce heat exchanger fouling, Proceeding of Heat Exchanger Fouling-Fundamental
Approaches and Technical Solutions, Davos, Switzerland, pp. 27-34.

Fowkes, F. M., 1964, Attractive forces at interfaces, Industrial and Engineering Chemistry,
Vol. 56, pp. 40-52.

Fox, H. W., Zisman, W. A., 1950, The spreading of liquids on low energy surfaces, I. PTFE,
Journal of Colloidal Science, Vol. 5, pp. 514-531.

Fritz, J. S., Schenk, G. H., 1987, Quantitative Analytical Chemistry, 5th ed., Allyn and Bacon
Inc., U.S.A.

Geddert, T., Bialuch, I., Augustin, W., Scholl, S., 2009, Extending the induction period of
crystallization fouling through surface coating, Heat Transfer Engineering, Vol. 30, pp. 868-
875.

Geddert, T., Kipp, S., Augustin, W., Scholl, S., 2007, Influence of different surface materials
on nucleation and crystal growth in heat exchangers, Proceeding of Heat Exchanger Fouling
and Cleaning: Challenges and Opportunities, Tomar, Portugal, Vol. RP5, pp. 229-236.

Ghaddar, N. K., Korczak, K., Mikic, B. B., Patera, A. T., 1986, Numerical investigation of
incompressible flow in grooved channels, Journal of Fluid Mechanics, Vol. 163, pp. 99-127.

Ginl, M., Sinn, G., Gindl, W., Reiterer, A., Tschegg, S., 2001, A comparison of different
methods to calculate the surface free energy of wood using contact angle measurements,
Colloids and Surfaces, A: Physicochemical and Engineering Aspects, Vol. 181, pp. 279–287.

116
List of References

Girifalco, L. A., Good, R. J., 1957, A theory for the estimation of surface and interfacial free
energy, I. derivation and application to interfacial tension, Journal of Physics and
Chemistry, Vol. 61, pp. 904-909.

Glater, J., York, J. L., Campbell, K. S., 1980, Scale formation and prevention in principles of
desalination. Editted by Spiegler, K. S., Laird A. D. K, Academic Press, New York, pp. 627-
650.

Gnielinski, V., 1978, Gleichungen zur Berechnung des Wärme- und Stoffaustausches in
durchströmten ruhenden Kugelschüttungen bei mittleren und grossen Pecletzahlen. VT-
Verfahrenstechnik, Vol. 12, pp. 363–366.

Good, R. J., Chaudhury, M. K., van Oss, C. J., 1991, Theory of adhesive forces across
interfaces 2. Interfacial hydrogen bonds as acid-base phenomena and as factors enhancing
adhesion. In: L.H. Lee Editor, Fundamentals of Adhesion, Plenum Press, New York, pp. 153-
172.

Good, R. J., 1992, Contact angle, wetting and adhesion, a critical review, Journal of Adhesion
Science Technology, Vol. 6, pp. 1269-1302.

Greiner, M., 1991, An experimental investigation of resistant heat transfer enhancement in


grooved channels, International Journal of Heat and Mass Transfer, Vol. 34, pp.1381-1391.

Gunn, D. J., 1980, Effect of surface roughness on the nucleation and growth of calcium
sulphate on metal surface, Journal of Crystal Growth, Vol. 50, pp. 533-540.

Hasson, D., Zahavi, J., 1970, Mechanism of calcium sulphate scale deposition on heat
transfer surfaces, Industrial and Engineering Chemistry Fundamentals, Vol. 9, pp. 1-10.

Heinzel, V., Jianu, A., Sauter, H., 2007, Strategies against particle fouling in the channels of
a micro heat exchanger when performing μPIV flow pattern measurements, Journal of Heat
Transfer Engineering, Vol. 28, pp. 222-229.

117
List of References

Israelachvili, J. N., 1991, Intermolecular and Surface Forces, 2nd ed., California, USA.

Karbowiak, T., Debeaufort, F., Voilley, A., 2006, Importance of surface tension
characterization for food, pharmaceutical and packaging products: a review, Critical Reviews
in Food Science and Nutrition, Vol. 46, pp. 391-407.

Kazi, N., Duffy, G., Chen, X. D., 2001, A study of fouling and fouling mitigation on smooth
and roughened metal surfaces, and a polymeric material, Proceeding of Heat Exchanger
Fouling: Fundamental Approaches and Technical Solutions, Davos, Switzerland, pp. 9-17.

Kazmierczat, T. F., Tomson, M. B., Nancollas, G. H., 1982, Crystal growth of calcium
carbonate, Journal of Physics and Chemistry, Vol. 86, pp. 103-107.

Kern, D. Q., Seaton, R. E., 1959, A theoretical analysis of thermal surface fouling, British
Chemical Engineering, Vol. 4 (5), pp. 258-262.

Keysar, S., Semiat, R., Hasson, D., Yahalom, J., 1994, Effect of surface roughness on
morphology of calcite crystallization on mild steel, Journal of Colloid and Interface Science
Vol. 162, pp. 311-319.

Kline, S. J., McClintock, F. A., 1953, Describing uncertainties in single-sample experiments,


Mechanical Engineering, Vol. 75, pp. 3-8.

Kukulka, D. J., Devgun, M., 2007, Fluid temperature and velocity effect on fouling, Applied
Thermal Engineering, Vol. 27 (16), pp. 2732-2744.

Kwang-Ho, C., Chung-Hak, L., 2000, Understanding membrane fouling in terms of surface
free energy changes, Journal of Colloid and Interface Science, Vol. 226, pp. 367-370.

Kwok, D. Y., Neumann, A. W., 1996, A simple experimental test of the Lifshitz-van der
Waals/acid-base approach to determine interfacial tensions, Canadian Journal of Chemical
Engineering, Vol. 74, pp. 551-553.

118
List of References

Lee, L. H., 1993, Roles of molecular interactions in adhesion, adsorption, contact angle, and
wettability, Journal of Adhesion Science and Technology, Vol. 7, pp. 583-634.

Liu, S., Nancollas, G. H., 1970, The kinetics of crystal growth of calcium sulphate dihydrate,
Journal of Crystal Growth, Vol. 6, pp. 281-289.

Liu, Y., Zhao, Q., 2005, Influence of surface energy of modified surfaces on bacterial
adhesion, Biophysical Chemistry, Vol. 117, pp.39-45.

Mahato, B. K., Shemilt, L. W., 1968, Effect of surface roughness on mass transfer, Chemical
Engineering Science, Vol. 23, pp. 183-185.

Malayeri, R. M., Al-Janabi, A., Müller-Steinhagen, H., 2009, Application of nano-modified


surfaces for fouling mitigation, International Journal of Energy Research, Vol. 33, pp. 1101-
1113.

Mantel, M., Wightman, J. P., 1994, Influence of the surface chemistry on the wettability of
stainless steel, Surface and Interface Analysis, Vol. 21, pp. 595-605.

Marshall, W., Slusher, R., Jones, E. V., 1964, Solubility and thermodynamic relationships for
CaSO4 in NaCl-H2O solutions from 40oC to 200oC, 0 to 4 Molal NaCl, Journal of Chemical
and Engineering Data, Vol. 9, pp. 187-191.

McGuire, J., Swartzel, K. R., 1989, The influence of solid surface energetics on
macromolecular adsorption from milk, Journal of Food Processing and Preservation, Vol.
13, pp. 145-160.

Mott, I. E. C., 1991, Biofouling and Studies Using Simulated Cooling Water, PhD
dissertation, University of Birmingham, Birmingham, U.K.

Müller-Steinhagen, H., Zhao, Q., 1997, Investigation of low fouling surface alloys made by
ion implantation technology, Chemical Engineering Science, Vol. 52, pp. 3321-3332.

119
List of References

Müller-Steinhagen, H., Zhao, Q., Reiss, M., 1997, Ion implantation-a new method of
preparing low fouling metal surfaces, Engineering Foundation Conference on Understanding
Heat Exchanger Fouling and Its Mitigation, Lucca, Italy.

Mullin J. W., 1993, Crystallization, 3th ed., Butterworth-Heinemann, Oxford.

Najibi, S., Müller-Steinhagen, H., Jamialahmadi, M., 1996, Calcium sulphate scale
formation during subcooled flow boiling, Chemical Engineering Science, Vol. 52 (8), pp.
1265-1284.

Neumann, A. W., Good, R. J., Hope, C. J., Sejpal, M., 1974, An equation-of-state approach
to determine surface tensions of low-energy solids from contact angles, Journal of Colloid
and Interface Science, Vol. 49, pp. 291-304.

Oliveira, R., 1997, Understanding adhesion: a means for preventing fouling, Experimental
Thermal and Fluid Science, Vol. 14, pp. 316-322.

Owens, D. K., Wendt, R. C., 1969, Estimation of surface free energy of polymers, Journal of
Applied Polymer Science, Vol. 13, pp. 1741-1747.

Packham, D. E., 2005, Handbook of Adhesion, 2nd ed., John Wiley and Sons Ltd., Chichester,
U.K.

Palache, C., Berman, H., Frondel, C., 1951, Dana’s System of Mineralogy, 7th ed., New York.

Panchal, C. B., Knudsen, J. G., 1998, Mitigation of water fouling: technology status and
challenges, Advances in Heat Transfer, Vol. 31, pp. 431– 474.

Patel, S., Finan, M. A., 1999, New antifoulants for deposit control in MSF and MED plants,
Desalination, Vol. 124, pp. 63-74.

Patridge, E. P., White, A. H., 1929, The solubility of calcium sulphate from 0 to 200, Journal
of Chemical Society, Vol. 51, pp. 360-370.

120
List of References

Rahman, M. M., Gui, F., 1993, Experimental measurements of fluid flow and heat transfer in
microchannel cooling passages in a chip substrate, Advances in Electronic Packaging, Vol. 4,
pp. 685-692.

Rankin, B. H., Adamson, W. L., 1973, Scale formation as related to evaporation surface
conditions, Desalination, Vol. 13, pp. 63-87.

Reegen, S. L., Ilkka, G. A., 1962, Adhesion and Cohesion, P. Weiss, Ed.; El-sevier: New
York.

Rizzo, G., 2008, Induktionszeit beim Kristallisationsfouling an ionenimplantierten


Wärmeübertragerflächen, PhD dissertation, ITW Institute, Stuttgart University, Germany.

Rosmaninho, R., Melo, L. F., 2006, Calcium phosphate deposition from simulated milk
ultrafiltrate on different stainless steel-based surfaces, International Dairy Journal, Vol. 16,
pp. 81–87.

Rosmaninho, R., Rizzo, G., Müller-Steinhagen, H., Melo, L. F., 2005, Anti-fouling stainless
steel based surfaces for milk heating processes. Proceeding of Heat Exchanger Fouling and
Cleaning: Challenges and Opportunities, Kloster Irsee, Germany, Vol. RP2, pp. 97-102.

Rosmaninho, R., Roch, F., Rizzo, G., Müller-Steinhagen, H., Melo, L. F., 2007, Calcium
phosphate fouling on TiN-coated stainless steel surfaces: role of ions and particles, Chemical
Engineering Science, Vol. 62, pp. 3821-3831.

Rosmaninho, R., 2007, Fouling Caused by Milk Components on Modified Surfaces during
Heat Treatment Processes, PhD dissertation, Faculty of Engineering of the University of
Porto, Portugal.

Santos, O., Nylander, T., Rizzo, G., Müller-Steinhagen, H., Trägårdh, C., Paulsson, M.,
2003, Study of whey protein adsorption under turbulent flow, Proceeding of Heat Exchanger
Fouling and Cleaning: Fundamentals and Applications, Santa Fe, New Mexico, USA, Vol.
RP1, pp. 175-183.

121
List of References

Sharma, P. K., Rao, K. H., 2002, Analysis of different approaches for evaluation of surface
energy of microbial cells by contact angle goniometry, Advances in Colloids and Interfaces
Science, Vol. 98, pp. 341-463.

Sharma, P. K., Rao, K. H., 2003, Adhesion of paenibacillus polymyxa on chalcopyrite and
pyrite: surface thermodynamics and extended DLVO theory, Colloids and Surfaces B, Vol.
29, pp. 21-38.

Strickland-Constable R. F., 1968, Kinetics and Mechanisms of Crystallization, Academic


Press, New York.

van Oss, C. J., Good, R. J., Chaundhury, M. K., 1986, The role of van der Waals forces and
hydrogen bonds in ‘hydrophobic interactions’ between biopolymers and low energy surfaces,
Journal of Colloid Interface Science, Vol. 111, pp. 378-390.

van Oss, C. J., Good, R. J., Chaundhury, M. K., 1987, Determination of the hydrophobic
interaction energy-application to separation processes, Separation Science and Technology,
Vol. 22, pp. 1-24.

van Oss, C. J., Good, R. J., Chaundhury, M. K., 1988, Additive and non additive surface-
tension components and the interpretation of contact angles, Langmuir 4, pp. 884-891.

van Oss, C.J., 1994, Interfacial Forces in Aqueous Media, Marcel Dekker, New York.

Verwey, E. J. W., Overbeek, J. Th. G., 1948, Theory of the Stability of Lyophobic Colloids.
Elsevier Pub. Co. Inc.

Visser, H., 2001, Improvement of construction materials used in the food industry to lengthen
processing time-a new European project (MODSTEEL), Proceeding of Heat Exchanger
Fouling: Fundamental Approaches and Technical Solutions, Davos, Switzerland, pp. 3-10.

Volmer, M., 1929, Über keimbildung und keimwirkung als specialfalle der heterogenen
katalyse, Z. Elektrochem, Vol. 35, p. 555.

122
List of References

Volmer, M., 1939, Kinetik der Phasenbildung, Steinkopf, Leipzig.

Wang, C. Y., 2003, Flow over a surface with parallel grooves, Physics of Fluids, Vol. 15, pp.
1114-1121.

Wang, G., Vanka, S. P., 1995, Convective heat transfer on periodic wavy passages,
International Journal of Heat and Mass Transfer, Vol. 38, pp. 3219-3230.

Wu, S., 1971, Calculation of interfacial tension in polymer systems, Journal of Polymer
Science Part C-Polymer Symposium, Vol. 34, pp. 19-30.

Wu, S., 1973, Polar and nonpolar interactions in adhesion, Journal of Adhesion, Vol. 5, pp.
39-55.

Wu, S., 1982, Polymer Interface and Adhesion, Marcel Dekker inc., New York.

Wu, S., 1999, in Polymer Handbook, 4th ed., Wiley, New York.

Wu, W., Du, J. H., Ma, B., Wang, B. X., 2002, Grooved wall effect on forced convective heat
transfer in a packed channels between two parallel plates, Journal of Enhanced Heat
Transfer, Vol. 9, pp. 117-121.

Wu, W., Nancollas, G. H., 1998, Kinetics of heterogeneous nucleation of calcium phosphate
on anatase and rutile surfaces, Journal of Colloids and Interface Science, Vol. 199, pp. 206-
211.

Wu, W., Zuang, H., Nancollas, G. H., 1997, Heterogeneous nucleation of calcium phosphate
on solid surfaces in aqueous solution, Journal of Biomedical Materials Research, Vol. 35, pp.
93-97.

Yoon, J., Lund, D. B., 1994, Magnetic treatment of milk and surface treatment of plate heat
exchangers: effects on milk fouling, Journal of Food Science, Vol. 59 (5), pp. 964-980.

123
List of References

Young, T., 1805, An essay on the cohesion of fluids, Philosophical Transactions of the Royal
Society, London, Vol. 95, pp. 65-87.

Zettler, H., 2002, Effect of Surface Properties and Flow Distribution on Fouling of Heat
Transfer Surfaces, PhD dissertation, Chemical and Process Engineering Department,
University of Surrey, Guildford, England.

Zettler, H., Weiß, M., Zhao, Q., Müller-Steinhagen, H., 2005, Influence of surface properties
and characteristics on fouling in plate heat exchangers, Heat Transfer Engineering, Vol. 26,
pp. 3-17.

Zhao, Q., Liu, Y., Müller-Steinhagen, H., Liu, G., 2002, Graded Ni–P–PTFE coatings and
their potential applications, Surface and Coating Technology, Vol. 155, pp. 279-284.

Zhao, Q., Müller-Steinhagen, H., 2002, Intermolecular and adhesion forces of deposits on
modified heat transfer surfaces, Proceeding of Heat Exchanger Fouling-Fundamental
Applications and Technical Solutions, Essen, Germany, pp. 41-46.

Zhao, Q., Liu, Y., Wang, C., Wang, S., Müller-Steinhagen, H., 2005a, Effect of surface free
energy on the adhesion of biofouling and crystalline fouling, Chemical Engineering Science,
Vol. 60, pp. 4858-4865.

Zhao, Q., Liu Y., Wang, S., 2005b, Surface modification of water treatment equipment for
reducing CaSO4 scale formation, Desalination, Vol. 180, pp. 133-138.

Zhao, Q., Xueju S., Wang, S., Xiaoling Z., Navabpour, P., Teer, D., 2009, Bacterial removal
properties of silicon and nitrogen-doped diamond-like carbon coatings, Biofouling, Vol. 25
(5), pp. 377-385.

Zisman, W. A., 1964, Relation of the equilibrium contact angle to liquid and solid
constitution, Advances in Chemistry, Vol. 43, pp. 1-51.

124
Appendixes

APPENDIXES

Appendix A1: Determination of the Surface Temperature


In this study, the heat is transmitted by thermal conduction through the specimen and
by forced convection in the flowing fluid as expressed by Eq. (A1).

1 x 1
= + (A1)
U k α

For given physical properties, the heat transfer coefficient (α) is directly proportional
to the Nusselt number as per Eq. (A2)

Nu k f
α= (A2)
D

In such a case, Eq. (A1) can be expressed as

1 x C
= + (A3)
U k Nu

The Nusselt number is calculated based on the following correlation (Gnielinski,


1978).

⎡ ⎛ D ⎞ 23 ⎤
Nu = 0.012 Re ( 0.87
)
− 280 Pr 0.4
⎢1 + ⎜ ⎟ ⎥ (A3)
⎢⎣ ⎝ L ⎠ ⎥⎦

The overall heat transfer coefficient is obtained for several liquid velocities from:

1 Tth − Tb
= (A4)
U q&

125
Appendixes

Afterwards, the results of 1/U are plotted against 1/Nu results. As expected, a linear
relationship (straight line) has been observed as shown in Figure A1. The intercept point of
this line represents the value of x/k. For the evaluation shown in Figure A1, x/k equals 0.0435
m2 K/kW.

measurement data
0.0005 linear fit

0.0004
1/U [m K/W]
2

y = 4.35E-5+0.0159x
0.0003

0.0002

0.005 0.010 0.015 0.020 0.025 0.030

1/Nu

Fig. A1 Determination of the x/k value between the heat transfer surface and the
thermocouples at position 1 (see Figure 3.3)

126
Appendixes

Appendix A2: Error Analysis

Consider “Y” as an objective function (here for instance fouling resistance) which has
to be calculated from a set of measurements as:

Y = f ( X 1 , X 2 , X 3 , ......., X M ) (B1)

The bias error of “Y” in turn is proportional to the partial gradient of “Y” to “Xi” and thus:

∂Y
BYi = Bi (B2)
∂X i

When several independent variables are involved then:

1/ 2
⎛ M ⎛ ∂Y ⎞ ⎞⎟
2

BY = ∑ ⎜⎜ B ⎟ (B3)
⎜ i =1 ⎝ ∂X i X i ⎟⎠ ⎟
⎝ ⎠

The precision error of any individual independent variable, Xi, is determined as the
standard deviation.

1/ 2
⎛ N
(
⎜ ∑ Xi − Xi )
2 ⎞

PYi = ⎜ i =1 ⎟ (B4)
⎜ N ( N − 1) ⎟
⎜ ⎟
⎝ ⎠

where “N” is the number of readings of each “X” for a given time. Like the bias error, the
overall precision error of “Y” can be determined as:

1/ 2
⎛ M ⎛ ∂Y ⎞ ⎞⎟
2

PY = ∑ ⎜⎜ P ⎟ (B5)
⎜ i =1 ⎝ ∂X i X i ⎟⎠ ⎟
⎝ ⎠

For the determination of uncertainty, the following information must be available:

127
Appendixes

• Bias error ( BYi ). In this study, it has been found that the bias error arises
from systematic experimental errors of i) approximately ± 0.2oC in
temperature measurement and ii) about ± 4.8% for the determination of the
heat flux as a result of systematic errors in the measurement of electrical
current and voltage.

• Mean value of a set of “N” observations of the measurement. According to


the guidelines of the ASME Journal of Heat Transfer Editorial Board for
estimating uncertainty, a sufficient number of samples (>30) should be
taken over a sufficient sampling period.

For a typical fouling experiment, the results are normally presented either as the
fouling resistance and/or the heat transfer coefficient. Figure A2 presents a flowchart for the
determination of uncertainty of fouling resistances.

⎛ ⎞
R f = f ⎜ q& , Tst , Tso ⎟
⎝ ⎠

2 2 2
⎛ ∂R f ⎞ ⎛ ∂R f ⎞ ⎛ ∂R f ⎞
Bias error (B) ⎜⎜ B q . ⎟⎟ ⎜ B ⎟ ⎜⎜ BTso ⎟⎟
⎝ ∂ q& ⎜ ∂T st ⎟
⎝ ∂Tso
T
⎠ ⎝ st ⎠ ⎠

2 2 2
Precision error (P) ⎛ ∂R f ⎞ ⎛ ∂R f ⎞ ⎛ ∂R f ⎞
⎜⎜ Pq ⎟⎟ ⎜⎜ PTst ⎟⎟ ⎜ P ⎟
⎜ ∂T so ⎟
⎝ ∂q& ⎝ ∂Tst
T
⎠ ⎠ ⎝ so ⎠

[
U 0.95 = B 2 + P 2 ]
1/ 2

Fig. A2 Bias and precision error terms for a typical fouling test

128
Appendixes

Appendix A3: Experimental Data by Förster (Förster, 2001)

Surface Surface energy components [mJ/m2] Experimental results


γ LW
2 γ +
2 γ −
2
initial fouling rate
[10-5 m2K/W h]
PFA 10.59 0.402 0 0.64
PTFE 11.26 0.2 0 1
FEP 13.94 0 0.52 0.836
Ms(I) 31.77 0.005 0.15 1.52
Cu 48 6.3 11.2 2.05
Ms(II) 33.9 0 0 0.565
Al 40.75 0.2 11 0.633
Bronze 43.7 0.001 15.25 0.621

129
Appendixes

Appendix A4: Experimental Data obtained by Zhao et al. (2009)

Coating Chemistry Surface energy components Adhered bacteria


name [mJ/m2] [CFU/cm2]
γ 2LW γ 2AB γ 2+ γ 2−
DLC1 1%Si-DLC 46.78 5.63 5.21 1.52 4237
DLC2 2%Si-DLC 47.43 5.56 0.96 8.05 4114.4
DLC3 3.8%Si-DLC 47.58 6.06 0.57 16.11 4359.7
DLC4 0.5%Si2.2%N-DLC 44.77 7.5 2.1 6.68 1485
DLC5 3.7%Si2.1%N-DLC 44.18 7.41 1.9 7.21 1471.4
DLC6 9.0%Si2.3%N-DLC 42.47 5.31 1.02 6.9 1171.7
DLC7 20%Si2.0%N-DLC 39.12 1.86 0.12 7.2 1566.8

130
List of Publications

LIST OF PUBLICATIONS
A) PEER-REVIEWED ARTICLES IN INTERNATIONAL JOURNALS

• Al-Janabi, A., Malayeri, R. M., Müller-Steinhagen, H., Minimization of CaSO4


Deposition through Surface Modification, Heat Transfer Engineering, Vol. 32, pp.
291-299 (2011).

• Al-Janabi, A., Malayeri, R. M., Müller-Steinhagen, H., Experimental Fouling


Investigation with Electroless Ni-P coatings, International Journal of Thermal
Sciences, Vol. 49, pp. 1063-1071 (2010).

• Malayeri, R. M., Al-Janabi, A., Müller-Steinhagen, H., Application of Nano-Modified


Surfaces for Fouling Mitigation, International Journal of Energy Research, Vol. 33,
pp. 1101-1113 (2009).

• Malayeri, M. R., Al-Janabi, A., Müller-Steinhagen, H., Impact of Surface Treatment


on the Fouling of Modified Heat Transfer Surfaces, Netsu Shori - Journal of the Japan
Society for Heat Treatment, Vol. 49, pp. 288-291 (2009).

• Al-Janabi, A., Malayeri, R. M., Müller-Steinhagen, H., Experimental Investigation of


Crystallization Fouling on Grooved Stainless Steel Surfaces during Convective Heat
Transfer, Heat Transfer Engineering (Special Issue), Vol. 30, pp. 832-839 (2009).

• Herz, A., Malayeri, M. R., Müller-Steinhagen, H., Fouling of Roughened Stainless


Steel Surfaces during Convective Heat Transfer to Aqueous Solutions, Energy
Conversion and Management (Special Issue), Vol. 49, pp. 3381-3386 (2008).

B) PEER-REVIEWED FULL ARTICLES IN PROCEEDINGS

• Al-Janabi, A., Esawy, M., Malayeri, M. R., Müller-Steinhagen, H., Consideration of


Dynamic Uncertainty in Fouling Experimentation, EUROTHERM Int. Conf. on Heat
Exchanger Fouling and Cleaning, Schladming, Austria, pp. 217-220, 2009.

• Al-Janabi, A., Malayeri, M. R., Müller-Steinhagen, H., Badran, O. O., Precipitation


Fouling on Various Austenitic Alloys, EUROTHERM Int. Conf. on Heat Exchanger Int.
Conf. on Heat Exchanger Fouling and Cleaning, Schladming, Austria, pp. 332-339,
2009.

131
List of Publications

• Al-Janabi, A., Malayeri, R. M., Müller-Steinhagen, H., Minimization of CaSO4


Deposition through Surface Modification, EUROTHERM Int. Conf. on Heat Exchanger
Int. Conf. on Heat Exchanger Fouling and Cleaning, Schladming, Austria, pp.289-296,
2009.

• Malayeri, M. R., Al-Janabi, A., Müller-Steinhagen, H., Fouling of Nano-Modified


Surfaces, the 4th International Conf. on Surfaces, Coatings and Nanostructured Materials
(NANOSMAT), Rome, Italy, 2009.

• Al-Janabi, A., Malayeri, M. R., Müller-Steinhagen, H., Meyer, F., Nonninger, R.,
Application of Environmentally-Friendly Organic/Inorganic Coatings for the Mitigation
of Crystallization Fouling, International Congress of Condition Monitoring and
Diagnostic Engineering Management (COMADEM), Prague, Czech Republic, 2008.
ISBN: 978-80-254-2276-2.

132
ISBN 978-3-00-035657-5

Das könnte Ihnen auch gefallen