You are on page 1of 10

Journal of Petroleum Science and Engineering 46 (2005) 185 194

www.elsevier.com/locate/petrol

Calcium carbonate scale formationassessing the initial


stages of precipitation and deposition
Tao Chena, Anne Nevillea,T, Mingdong Yuanb
a

Corrosion and Surface Engineering Research Group, School of Engineering and Physical Sciences, Heriot-Watt University,
Riccarton, Edinburgh, EH14 4AS, UK
b
Baker Petrolite, 12645 W. Airport Boulevard, Sugarland, Texas 77478, USA
Received 5 November 2003; accepted 7 December 2004

Abstract
Scale formation is a serious problem encountered in many industries including oil or gas production, water transport, power
generation and batch precipitation. Normally, studies of scale formation have been focused on precipitation processes in the
bulk solution using bulk jar methods where the precipitation tendency rate and inhibitor effectiveness are quantified. Several
recent studies have started to focus on scale deposits formed on the surface of metals. In this paper, calcareous scale formation
was studied both in the bulk solution and on the metal surface in three supersaturated scale formation solutions which represent
typical waters encountered in oil and gas production. An electrochemical technique, using a rotating disk electrode (RDE), was
used to quantify scale formation on the metal surface. With this technique, reduction of oxygen was considered at the surface of
a RDE. The rate of oxygen-reduction at the surface of the RDE enables the extent of surface coverage of scale to be assessed. To
understand the formation and growth of the surface scale deposit, surface analysis was used in conjunction with this technique.
Scanning electron microscopy (SEM) was used for analyzing the microstructure of the scale. At the same time, inductively
coupled plasma (ICP) was used for analyzing the quantity of the precipitate formed in the bulk solution and scale formed on the
metal surface by dissolving the scale.
In this paper it is demonstrated that bulk precipitation and surface deposition have different dependencies on the index of
supersaturation and so to completely understand an industrial scaling system both processes should be studied.
D 2005 Elsevier B.V. All rights reserved.
Keywords: Scale formation; Homogeneous precipitation; Heterogeneous precipitation; Calcium carbonate; Rotating disk electrode

1. Introduction

* Corresponding author. Fax: +44 131 451 3129.


E-mail address: a.neville@leeds.ac.uk (A. Neville).
0920-4105/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.petrol.2004.12.004

The formation of mineral scale, in particular


calcareous deposits, is a persistent and expensive
problem in industries ranging from oil and gas to
desalination. Scaling of metallic or insulating walls in

186

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

contact with water supersaturated with respect to


calcium carbonate may create technical problems
including impedance of heat transfer, increase of
energy consumption and unscheduled equipment
shutdown (Klepetsanis et al., 1995; Kjellin et al.,
2001; Qingfeng, 2001).
The precipitation of calcium carbonate has been
widely studied (Sohnel and Mullin, 1982; Abtahi et
al., 1996; Nancollas and Reddy, 1973). Traditionally,
studies of scale formation have concentrated on
assessing precipitation formed in the bulk solution
by using laboratory beaker or bulk jar tests (Abtahi et
al., 1996; Nancollas and Reddy, 1973). The research
has primarily focused on the assessment of the
kinetics of homogenous and heterogeneous precipitation in the bulk solution (Langelier, 1936). The
formation of the solid phase is throughout the mother
phase without any foreign solid phase in homogenous
processes, and the formation of new solid phase
particles is catalyzed by the presence of a foreign solid
phase in heterogeneous processes (Sohnel and Garside, 1992).
It has been demonstrated (Harris and Marshall,
1981) that there are often wide anomalies between
actual deposition on component surfaces and scaling
rates estimated by predictive models based on scaling
indices and thermodynamics to predict precipitation
tendency (Hasson et al., 1996). However, the relationship between precipitation and scale deposition on
solid surfaces has received little attention.
Some focus has been turned to this aspect of
scaling and has resulted in numerous studies
reporting methods to detect and assess scale
formation on metal surfaces. Hasson (Hasson et
al., 1996) and Zhang (Zhang et al., 2001), studied
calcium carbonate scale formation in a pipe flow
system in oil and gas industry and desalination
industry to attempt to overcome some of the
shortfall of beaker tests. Sullivan (Sullivan et al.,
1996) studied scale formation by monitoring the
heat transfer change. These methods have been
developed primarily to assess efficiency of scale
inhibitors and give only a relative estimate of the
thickness of scale. They are usually not sensitive
enough to study the primary layer of scale. In recent
work by the authorsT studies of scale formation on
metal surfaces have been progressed. Through
directing more attention to the scale formation and

adhesion at metal electrodes, in parallel with


precipitation processes the entire scaling process
can be characterized. This will then facilitate the
development of inhibitors where the functionality
can be targeted to the particular scaling problem.
In the work reported herein, an electrochemically
based technique is used to develop an understanding
of the kinetics of scale deposition on metal electrodes. The technique developed by Neville et al.
(1999) can quantify the scale coverage in the early
stages of scaling. In this paper comparisons are then
made with the bulk precipitation processed occurring
simultaneously.

2. Experiment techniques
2.1. Reagents
To create the supersaturated solutions for this
study, calcium chloride (CaCl2d 6H2O), sodium chloride (NaCl), magnesium chloride (MgCl2d 6H2O),
sodium bicarbonate (NaHCO3) of analytical grade
supplied by BDH Laboratory were used. The solutions were made using distilled water.
2.2. Procedures and measurements
Experiments were carried out in a 1-l vessel
thermostated at 20 8C. 800-ml scale formation
solution was used in each experiment. CaCO3 was
precipitated spontaneously by mixing two solutions
(400 ml brine 1 containing calcium ions and 400 ml
brine 2 containing bicarbonate ions). The pH was
buffered to 6.7 by acetic acid. Before mixing, the two
solutions were filtered using a 0.45-Am filter. Threescale formation solutions, the composition provided
by worker in the soil and gas industry, were used:
solutions A, B and C, respectively. S is the supersaturation ratio of these solutions. For CaCO3, S can
be expressed as:


S aCa2 aCO2
=Ksp
3
where a i is the activity of a given ion and K sp is the
solubility product of the scale forming mineral (here
CaCO3). The supersaturation index (SI), equal to log
S, is also commonly used to express scaling tendency.

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194


Table 1
Composition and SI of the scale formation solutions
Ions
Na+
Mg2+
Ca2+
HCO
3
Cl
SI (=log S)

Solution A

Solution B

Solution C

Conc. (mg/l)

Conc. (mg/l)

Conc. (mg/l)

6873
91
1440
2196
12,136
1.25

7085
91
720
930
12,136
0.59

7085
91
350
533.75
12,136
0.04

The SI used in this study was calculate by ScaleSoftPitzerTM software provided by Rice University. The
composition and SI for each solution are given in
Table 1.
Calcium carbonate scales were deposited from the
scale formation solution to the active surface of a
stainless steel (UNS S31603) electrode, shown in
Fig. 1. The electrodes are referred to as Rotating
Disk Electrodes (RDE). Analysis of scale required
that they be rotated and exposed to a controlled
hydrodynamic regime. Before immersing in the scale
formation solution, the RDEs were abraded with
silicon carbide paper to 1200 grit and polished with
diamond polishing compound (6 Am). After the final
polishing, the specimens were rinsed with distilled
water and acetone and dried with compressed air. Six
RDE samples were immersed in the scale formation
solution and removed at 1, 4, 8, 12, 16 and 24 h. A
magnetic stirrer was used to stir the solution at a rate
of 300 rpm during the experiment. During the
deposition step, the RDEs were static.

187

To monitor the coverage of the scale on the metal


surface, electrochemical measurements on the RDE
were used, which have been presented previously
(Neville et al., 1999). The technique is based on the
measurement of the rate of oxygen-reduction on the
RDE under potentiostatic control. As scale is formed
on the electrode the active surface for oxygenreduction is reduced and this can be quantified by
measurement of the oxygen-reduction reaction. The
oxygen-reduction reaction is controlled by a mass
transfer in the conditions used in the study. The
limiting current of the oxygen reduction reaction is
related to the rotational speed of the RDE by the
Levich formula as shown in the following equation
(Adams, 1969):
iL 0:62nFACb D2=3 m1=6 x1=2
where n is the number of the electrons, F is
Faradays constant, A is surface area of the electrode
(cm2), t is kinematical viscosity of the electrolyte
(cm2 s1), C b is concentration of oxygen in the
solution (mol dm3), D is diffusion coefficient of
oxygen in the electrolyte (cm2 s1) and x is angular
velocity of the disc (rad s1).
The linear relationship between the limiting current
(i L) and square root of the rotational speed (x 1/2) has
been widely used in the determination of diffusion
coefficients or electro-active species concentration in
mass transport controlled systems (Filinovsky and
Pleskov, 1983). In this study, this linear relationship
has been used to detect the coverage of scale formed
on the metal surface.

Fig. 1. RDE sample for scale formation on the active surface of electrode.

188

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

The RDE system is shown in Fig. 2. The computer


drives the potentiostat and records current and
electrochemical parameters in the 3-electrode cell as
a function of time via the software Corrware. The
RDE system is made of a frame supporting a DCmotor, the speed sensor, the reference electrode (RE)
and the auxiliary electrode (AE). The motor is
connected to the working electrode (WE) via a
bearing and a chuck. The level of the beaker inserted
inside the frame is adjusted to allow the immersion of
the three electrodes.
The potentiostatic analysis procedure, which
allows the plot of i L versus x 1/2 to be assessed, is
constituted of an initial polarization of the RDE
sample at 0.8 V (Ag/AgCl). Reduction of oxygen
was considered at the surface of the RDE and
analysis of the sample by the electrochemical
technique involves recording the limiting current at
the RDE versus the rotational speed at a constant
potential of 0.8 V (Ag/AgCl). The range of
rotational speeds used is 6002200 rpm. The electrochemical analysis before and after scaling was
carried out in an analysis solution, which comprised
of 5 g/l sodium chloride solution, buffered at pH 10
by addition of NaOH prior to analysis. No scaling
constituents were present to ensure that no further
scaling occurred during this quantification step. In
order to check the coverage of the scale on the
working electrode, four steps are involved: (1) initial
analysis of the non-scaled working electrode by the
analysis of i L versus x 1/2; (2) immersion of the

sample electrode in a supersaturated solution of


calcium carbonate for a predetermined time; (3) final
i L versus x 1/2 analysis of the scaled working electrode
after a fixed time in the supersaturated solution; (4)
calculate the coverage.
The electrochemical technique described here to
quantify the extent of scaling relies on the distinction
between the active surface area on a bare and scaled
electrode. Fig. 3 is a simple representation of discrete
crystals on a surface where A i is the total active
surface area of the bare electrode and A f is the reduced
active surface area (the surface area covered by the
scale crystals) once scale crystals have deposited.
Before scaling, this active surface is equal to the entire
RDE surface which is uniformly accessible to O2reduction. The scale formed on the metal surface
reduces the active surface for the reduction of
dissolved oxygen.
Using the Levich analysis, the linear relationship
between the limiting current (i L) and square root of
the rotational speed (x 1/2) for the unscaled sample and
scaled sample is shown in Fig. 4. K i and K f are the
slopes of the i L versus x 1/2 lines, which are proportional to the area of active surface of the RDE. The
coverage of scale formed on the metal surface can be
expressed as:
%coverage

Ki  Kf
Ki

where K i and K f are the slopes of Levich analysis


before and after the deposition respectively. Validation

Fig. 2. Schematics of RDE system.

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

189

Fig. 3. Change of active surface of RDE before and after scale formation for oxygen-reduction.

of the %Coverage measured by electrochemical analysis


and by image analysis has been presented in a
previous communication (Neville et al., 1999).
Scanning electron microscopy (SEM) and light
microscopy were used for analyzing the scale
morphology.
The scale on the metal surface was dissolved by
10 ml ethylenediaminetetraacetic acid (EDTA) and
NaOH solution (10 g EDTA and 10 g NaOH to 1 l
solution) and then analyzed by Inductively Coupled
Plasma (ICP) of wavelength 317.933 nm to measure
the concentration of calcium ion. Hence the quantity
of calcium carbonate surface deposit can be calculated. Also, ICP is used for monitoring the calcium
ion concentration in bulk solution during the experiments. 10 ml bulk solution was removed from scaling
system and filtered by 0.75-Am filter paper naturally. 1
ml filtered bulk solution was diluted to 20 ml by

quenching solution composed of 1000 ppm active


polyvinyl sulfonate and polyacrylic acid copolymer
(PVS) and 3000 ppm KCl buffered to pH 8 to 8.5 by
0.1 M NaOH for ICP analysis to get the concentration
of calcium ions in the bulk solution. Hence the extent
of calcium carbonate bulk precipitation is calculated.

3. Results and discussion


Two aspects of CaCO3 scale formation will be
addressed in this study: precipitation in the bulk
solution and deposition onto a metal surface.
3.1. Bulk precipitation
Calcium carbonate precipitation in the bulk solution over a period of 24 h is shown in Fig. 5. It is

1000
900

Before deposition, Ki

800

iL (A)

700
600
500

After deposition, Kf

400
300
200
100
0
0

10

20

30

40

50

1/2 (rpm1/2)
Fig. 4. Limit current against square root of rotating speed for Levich analysis.

60

190

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

1400
1200
1000
800
600
400
200
0
0

10

15

20

25

time (hour)
A

Fig. 5. Precipitation in the bulk solution for solutions A, B and C measured by ICP analysis of the Ca ion over a 24-h period.

supersaturation index during the experiment when


precipitation occurs. Hence the driving force for
precipitation reduces as time elapses.

clear that calcium carbonate precipitation in the bulk


solution proceeds over the test period in solutions A
and B. In solution C, there is almost no precipitation
detected by ICP before 12 h. The experimental
induction period of precipitation is defined as the
time which elapses between the creation of supersaturation and the first observable change in some
physical property of the precipitating system, e.g. the
appearance of crystals or turbidity, change of solution
conductivity, change in solution composition, etc.
(Sohnel and Mullin, 1982). In this work, the induction
time in solution C is longer than 12 h and shorter than
1 h in solutions A and B. In solution A, the amount of
calcium precipitation increases rapidly in the first 8 h
and then a slower rate of scale precipitation occurs. It
is presumed that this is caused by the decrease of the

3.2. Surface scaling


Fig. 6 shows the surface coverage of the RDE
samples over the same period in scale formation
solutions A, B and C. The coverage increases as time
elapses in the three scale formation solutions. The
ranking of the coverage is: ANBNC, as expected. The
surface coverage is similar in solutions B and C. The
surface coverage reaches steady state after 4 h in
solution A and after 12 h in solutions B and C.
The quantity of CaCO3 scale formed on the metal
surface, as measured by dissolution of scale EDTA

120

coverage (%)

100
80
60
40
20
0
0

10

15

20

25

time (hour)
A

Fig. 6. Scale coverage on the metal surface in three scale formation solutions.

30

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

in solution B, large particles can be seen. There are no


large particles formed in solution C before 12 h.
However, small particles (sub-micron size) are clear,
distributed over the surface. In the view of the size
and the number of the crystals, the results can be
linked with the quantity of the scale formed on the
metal surface.
The crystals formed on the metal surface and in the
bulk solution at 24 h in solution A are shown in Fig. 9.
The size of the crystals on the metal surface is
between 10 and 20 Am. In the bulk solution, the size
of the crystals forming in the bulk precipitate is about
5 Am. It shows the precipitations formed in the bulk
are different with deposits formed on the metal surface
which is in agreement with result made by other
authors (Hasson et al., 1996).
In contrast to traditional studies which have
considered relatively thick layers of scale formed on
the metal surface, in this study the primary layer of
scale formed on metal surface was the focus. It
included not only the crystal growth period, but also
the crystal nucleation period. Calcium carbonate
crystallisation consists of two stages: nucleation and
growth. In the nucleation period, many small particles
form and the number of the particles increases very
quickly during this period. The growth of the crystals
takes place on the particles formed in the nucleation
time and the size of the particles increases at this
period. Pernot (Pernot et al., 1998) and others
(Koutsoukow and Kontoyannis, 1994) concluded that

and ICP analysis, is shown in Fig. 7. The calcareous


scale increases with time in the scale formation
solutions as expected from the coverage results. Also
similar scale quantities in solutions B and C are
recorded. It is apparent from Fig. 5 that there is no
precipitation in the bulk solution before 12 h.
However, it is evident there is a small amount of
scale formed on the metal surface during this time
both from the coverage analysis in Fig. 6 and quantity
analysis in Fig. 7.
Simultaneously, the coverage and the quantity of
deposit formed on the metal surface (Figs. 6 and 7)
show no great differences between solutions B and C.
However, the difference of the quantity of precipitation formed in the bulk solution is obvious (Fig. 5). It
shows that the deposit formed on the metal surface
and precipitation formed in the bulk solution, being
two different processes, each has their own mechanisms and kinetics.
3.3. Scale morphology
Fig. 8 shows light microscope images of the scale
formation on the metal surface. The images show a
regular distribution of crystals and are useful to get a
visible assessment of the induction time for the
surface deposition. Some large particles are formed
quickly on the surface in solution A and are visible at
1-h. This explains the non-zero readings from ICP
after dissolution of scale, presented in Fig. 7. From 4 h
0.35

quantity (g)

0.3
0.25
0.2
0.15
0.1
0.05
0

191

10

15

20

25

30

time (hour)

Fig. 7. Quantity of calcium carbonate scale on the metal (UNS S31603) surface in solutions A, B and C.

192

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

solution A: 1 hour

solution B: 1 hour

solution C: 1 hour

solution A: 4 hours

solution B: 4 hours

solution C: 4 hours

solution A: 12 hours

solution B: 12 hours

solution C: 12 hours

Fig. 8. Scale formed in solutions A, B and C at 1, 4 and 12 h metal surface bulk solution.

the number of precipitating calcium carbonate crystals


increases rapidly at the start of nucleation and at a
later stage the particles stop increasing in number and
subsequently grow. In Fig. 5, there is almost no
precipitation before 12 h in solution C, which was
regarded as the induction period in the bulk solution.
In this period, only some sub-micron size particles are
formed on the metal surface in solution C, as shown in
Fig. 8. In Fig. 6, the coverage on the metal surface at
12 h is up to 70%, which shows that the sub-micron
size particles are the main contribution to the coverage

metal surface

on the metal surface. In solution A at 4 h, the coverage


is about 85% (Fig. 6). Only a small number of large
particles were observed at this time (Fig. 8) and they
are not the main contribution to the coverage. Hence,
sub-micron size particles formed in the nucleation
period are the main contribution to the coverage in the
initial stages of surface scale formation.
The driving force for scale formation is supersaturation ratio (S) of the scale formation solution
which affects the induction time, growth, morphology
of crystals and rate of scale formation (Nancollas,

bulk solution

Fig. 9. Crystals formed on metal surface and in the bulk solution at 24 h in solution A.

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

1968; Turner and Smith, 1998; Jimene-Lopez et al.,


2001; Khan et al., 2001). In terms of thermodynamics,
three possibilities exist in terms of scale formation
from the solution. (1) Sb1: the solution is undersaturated and scale formation is not thermodynamically feasible. (2) S=1: the solution is saturated. The
scale formation and dissolution rate in the solution is
the same and no scale is formed in the solution. (3)
SN1: the solution is supersaturated and scale formation is thermodynamically possible (Legrand, 1990).
High supersaturation scale formation solutions tend to
promote precipitation in the bulk solution and scale
formed on the metal surface (Nancollas, 1968).
Fig. 6 showed that the high supersaturation scale
formation solution promoted more extensive coverage
of the scale formed on metal surface, especially at 1 h
of scale formation, in agreement with conclusions
made by other authors (Nancollas, 1968; Chen et al.,
1997; Zhang et al., 2001). The morphology of the
scale formed on the metal surface was shown in Fig.
8. The high supersaturation scale formation solution
accelerated crystal nucleation and growth (Chen et al.,
1997). In solution C, the supersaturation index is 1.10,
which is near 1. There was no precipitate observed
before 12 h in the bulk solution. However, scale was
formed on the surface of the metal. The effect of the
supersaturation ratio on the scale formed in the bulk
and on the metal surfaces in the solution A, B and C is
summarized as following: Ashort bulk induction
time and rapid growth. Short surface induction time
and rapid growth; Bshort bulk induction time but
slow growth. Short surface induction time still rapid
growth of coverage; Clong bulk induction time.
Short surface induction time still rapid growth of
coverage.
In this study, calcium carbonate precipitation in
the bulk solution is regarded as primarily a
homogenous process because the bulk solution is
filtered before the experiment, and calcium carbonate formed on the metal surface is primarily a
heterogeneous process. In this paper, the metal
surface is the foreign solid used for the deposit
formation in contrast to other studies using seeded
crystals like calcite crystals, sand particles, glass
particles (Sohnel and Garside, 1992; Vetter, 1980,
1987). In solution C, the induction time for the
precipitation in the bulk solution is longer than 12 h
(Fig. 5). In Figs. 5 and 6, there were some sub-

193

micron size crystals formed on the metal surface, the


quantity was detected by ICP at 1 h and the
coverage was up to 20% by RDE analysis, which
showed the nucleation time for the scale formation
on the metal surface in solution C is less than 1 h. It
is demonstrated that the induction time for the scale
formed on metal surface is less than precipitation
formed in the bulk solution confirming that heterogeneous crystal nucleation is easier than homogenous crystal nucleation (Sohnel and Garside, 1992).
In Fig. 9, the average size of the scale crystals
formed on the metal surface was larger than the
precipitation crystals formed in the bulk solution,
which implied that heterogeneous scale formation
conditions promoted crystal growth compared with
homogenous conditions (Sohnel and Garside, 1992).
Scale formed on the metal surface and precipitation
formed in the bulk solution are two different processes
and have different mechanisms of crystal nucleation
and growth. Since surface scale formation is normally
the process which causes operational problems in
industrial situations. It therefore warrants study along
with bulk precipitation. This is particularly important
when inhibition is considered. It is therefore important
and useful to fully study the scale formation directly
on the metal surface to understand real processes of
scale formation in industry.

4. Conclusions
! The supersaturation ratio of the scale formation
solution has an important effect on the induction,
growth, morphology of crystals and rate of scale
formation. In three different supersaturation ratio
solutions A, B and C: Ashort bulk induction
time and rapid growth. Short surface induction
time and rapid growth; Bshort bulk induction
time but slow growth. Short surface induction time
still rapid growth of coverage; Clong bulk
induction time. Short surface induction time still
rapid growth of coverage.
! Sub-micron size crystals are the main contribution
to coverage on the metal surface in the initial
stages of scale formation.
! The crystals of scale formed on the metal surface
are larger than the precipitated crystals formed in
the bulk solution, which confirms that heteroge-

194

T. Chen et al. / Journal of Petroleum Science and Engineering 46 (2005) 185194

neous conditions promote crystal growth compared


with homogeneous conditions.
! Studying both bulk precipitation and surface
scaling is necessary to completely understand real
industrial scaling processes.

Acknowledgements
The authors would like to thank Baker Petrolite
USA for funding this work and for the permission to
publish this paper.

References
Abtahi, Manoochehr, Kaada, Baard, Vindstad, Jens E., Qstvold,
Terje, 1996. Calcium carbonate precipitation and pH variations
in oil field waters. A comparison between experimental data and
model calculations. Acta Chemica Scandinavica 50, 114 121.
Adams, R., 1969. Electrochemistry of Solid Electrodes. Decker,
p. 11.
Chen, Pao-Chin, Tai, C.Y., Lee, K.C., 1997. Morphology and
growth rate of calcium carbonate crystals in a gasliquidsolid
reactive crystallizer. Chemical Engineering Science 52 (21/22),
4171 4177.
Filinovsky, V., Pleskov, Y.V., 1983. Eletrodiscs: experimental techniques. In: Yeager, E., Bockris, J. (Eds.), Comprehensive Treatise
of Electrochemistry. Plenum Press, New York, pp. 293 352.
Harris, A., Marshall, A., 1981. The evaluation of scale control
additives. Proceedings of Symposium on Progress in the
Prevention of Fouling in Industrial Plant, pp. 174 199.
Nottingham, April.
Hasson, D., Bramsom, D., Limoni-Relis, B., Semiat, R., 1996.
Influence of the flow system on the inhibitor action of CaCO3
scale prevention additives. Desalination 108, 67 79.
Jimene-Lopez, C., Caballero, E., Huertas, F.J., Romanek, C.S.,
2001. Chemical, mineralogical and isotope behaviour, and phase
transformation during the precipitation of calcium carbonate
minerals form intermediate ionic solution at 250 8C. Geochimica et Cosmochimica Acta 65 (19), 3219 3231.
Khan, M.S., Budair, M.O., Zubair, S.M., 2001. A parametric study
of CaCO3 scaling in AlSl 316 stainless steel tubes. Heat and
Mass Transfer 38, 115 121.
Kjellin, Per, Holmberg, Krister, Nyden, Magnus, 2001. A new
method for the study of calcium carbonate growth. Colloids and
Surfaces, A: Physicochemical and Engineering Aspects 194,
49 55.

Klepetsanis, Pavlos G., Lampeas, Nikolaos, Kioupis, Nikolaos,


Koutsoukos, Petros G., 1995. The effect of mineral deposits on
stainless steel. Mineral scale formation and inhibition. Zahid
Amjad. Plenum press, New York, pp. 131 144.
Koutsoukow, P., Kontoyannis, C., 1994. Precipitation of calcium
carbonate in aqueous solutions. Journal of the Chemical Society.
Faraday Transactions 1567, 17 21.
Langelier, W.F., 1936. The analytical control of the anti-corrosion
water treatment. Journal of the American Water Works
Association 28, 1500 1506.
Legrand, L., 1990. Presentation de la Corrosion et. de lentartrage
dans les metaux de Distribution de leau. Ellis Horwood Edition,
p. 30.
Nancollas, George H., 1968. Kinetics of crystal growth from
solution. Journal of Crystal Growth 3/4, 335 339.
Nancollas, G.H., Reddy, M.M., 1973. The kinetics of crystallization
of scale-forming minerals. SPE 4360, SPE-AIME oilfield
chemistry symposium, May 2425, pp. 117 126.
Neville, A., Hodgkiess, T., Morizot, A.P., 1999. Electrochemical
assessment of calcium carbonate deposition using a rotating disk
electrode (RDE). Journal of Applied Electrochemistry 29 (4),
455 462.
Pernot, B., Euvrard, M., Simon, P., 1998. Effects of iron and
manganese on the scaling potentiality of water. Journal of Water
Supply: Research and Technology. AQUA 47 (1), 21 29.
Qingfeng, Yang, 2001. Investigation of calcium carbonate scaling
inhibition and scale morphology by AFM. Journal of Colloid
and Interface Science 240, 608 621.
Sohnel, O., Garside, J., 1992. PrecipitationBasic Principles and
Industrial Applications. Butterworth Heinemann, Oxford.
Sohnel, O., Mullin, J.W., 1982. Precipitation of calcium carbonate.
Journal of Crystal Growth 60, 239 250.
Sullivan, PJ., Young, T., Carey, J., 1996. Effectiveness of polymer
phosphonate blends for inhibition of CaCO3 scale. Industrial
Water Treatment 26, 39 44 (November/December).
Turner, Carl W., Smith, David W., 1998. Calcium carbonate
scaling kinetics determined from radiotracer experiments with
calcium-47. Industrial & Engineering Chemistry Research 37,
439 448.
Vetter, O.J., 1980. Prediction of CaCO3 scale under downhole
conditions. SPE 8991, 5th Oilfield Geothermal Chemistry, CA
May, pp. 28 30.
Vetter, O.J., 1987. Calcium carbonate scale in oilfield operations.
SPE 16908, 62nd Ann. Tech. Conf., Dallas, TX. Sep. 2730,
pp. 27 30.
Zhang, Yuping, Shaw, H., Farquhar, R., Dawe, R., 2001. The
kinetics of carbonate scaling-application for the prediction of
downhole carbonate scaling. Journal of Petroleum Science and
Engineering 29, 85 95.

You might also like