Sei sulla pagina 1di 28

Int. J. Oil, Gas and Coal Technology, Vol. 1, Nos.

1/2, 2008 81

Mechanisms of scale deposition and scale removal


in porous media

Mohammad Jamialahmadi
The University of Petroleum Industry,
Ahwaz, Iran
E-mail: Jami@put.ac.ir

Hans Müller-Steinhagen*
Institute for Thermodynamics and Thermal Engineering,
University of Stuttgart,
Stuttgart, Germany

Institute of Technical Thermodynamics,


German Aerospace Center,
Stuttgart, Germany
E-mail: Hans.Mueller-Steinhagen@dlr.de
*Corresponding author

Abstract: Calcium sulfate scaling and scale removal have been studied
experimentally and theoretically in a well-defined unconsolidated porous
medium. The effects of injection rate, scale concentration and temperature on
the rate of scale formation and the rate of its removal are investigated. The
results of the experiments are used to develop mechanistic models for
the prediction of the scaling process in porous media and for removal of the
deposits by dilute sulfuric acid solutions. The predictions of the suggested
models for scaling and cleaning are compared with measured data. Deviations
ranging from 11% to 14% are observed, which confirm the suitability of the
models.

Keywords: scaling; porous media; calcium sulphate; porosity; permeability;


cleaning.

Reference to this paper should be made as follows: Jamialahmadi, M. and


Müller-Steinhagen, H. (2008) ‘Mechanisms of scale deposition and scale
removal in porous media’, Int. J. Oil, Gas and Coal Technology, Vol. 1,
Nos. 1/2, pp.81–108.

Biographical notes: Mohammad Jamialahmadi is a Professor of Chemical and


Petroleum Engineering at the University of Petroleum Industry, Ahwaz, Iran.
He received his BSc (1978), MSc (1982) and PhD (1982) in Chemical Process
Engineering from Aston University in Birmingham, UK. His research work
covers topics related to transport phenomena, fouling, multiphase flow, heat
and mass transfer in porous media, oil field formation damage due to particle
movement, asphaltene deposition and scaling, enhanced oil recovery and neural
networks. He has published more than 360 papers in international journals and
proceeding conferences. His work has been honoured with several

Copyright © 2008 Inderscience Enterprises Ltd.


82 M. Jamialahmadi and H. Müller-Steinhagen

international and national awards, including the Award from his Excellency
Mr. Khatami, President of Iran, for outstanding services to the National Iranian
Oil Co. (1998) and the International Kharazmi Award for fundamental research
in engineering (2001). He is a Member of the editorial board of the Iranian
Journal of Chemical Engineering, Petroleum Engineering and the Journal of
Scientia Iranica.

Hans Müller-Steinhagen is the Director of the Institute for Thermodynamics


and Thermal Engineering of the University of Stuttgart and Director
of the Institute of Technical Thermodynamics of the German
Aerospace Centre (DLR). He holds doctoral degrees from the Universities
of Karlsruhe, Auckland and Erlangen-Nürnberg. His research work
cover a wide range of topics related to heat and mass transfer, multiphase
flow, fuel cells, solar technology, process thermodynamics and
systems analysis. He is the author/editor of more than 450 books
and articles, for which he has received numerous national and international
awards. He is a Fellow of the Royal Academy of Engineering, a Fellow
of the Institution of Chemical Engineers and a UK Chartered Engineer.
He is on the editorial board of several international journals, as well as being
President of EUROTHERM and Member of the executive boards of EUREC
and ICHMT.

1 Introduction

Brackish water has been involved with the petroleum industry since the days of oil being
discovered. Almost all oil production wells produce some brackish water, but in some
cases, it may be small, and in other instances, as much as 90% of the total fluid flow.
If temperature, pressure, concentration of various species of water in the oil-bearing
formation, or surface conditions are changed naturally or deliberately, various inorganic
salts may precipitate in a certain order. Scale deposits are formed from components with
limited solubility in water, and in most cases increase with increasing temperature.
Species of dissolved salts in water, which may cause severe scaling problems
are the various forms of calcium sulphate, namely gypsum, anhydrate and hemi-hydrate.
Strontium and barium ions may also be present in the formation water. If sulphate rich
water such as sea water is injected to maintain pressure or displace the oil, these ions
may precipitate as their sulphate in the formation matrix and cause scaling problems.
Furthermore, the injected water may contain carbon dioxide under pressure, which can
lead to calcium carbonate scaling as the pressure decreases. Scaling in the petroleum
industry is categorised among the most critical and expensive problems, which may
happen in the reservoir, well tubing and surface facilities. Scaling problems generally
begin with drilling of the oil wells in drilling pipe, heating equipment, down-hole pumps,
tubing, casing and other drilling and production facilities. Scaling of safety valves can be
so severe that it may be impossible to close them properly when needed. Costs due to oil
field scaling and cleaning are generally high because they may be associated with a
drastic decline of oil production and increasing operating and down-time costs.
Exact cost figures are not available as in most oil companies, scaling costs are lumped
under broader categories like general maintenance costs. Scaling causes a large number
Mechanisms of scale deposition and scale removal in porous media 83

of costly operations such as perforation, redrilling, stimulation and other remedial work
in producing wells and in salt water disposal wells. It is estimated that the productivity of
oil wells in the USA has declined significantly because of scaling of formation matrices
and well tubing. It is not unlikely that many wells or even reservoirs are abandoned
permanently because of scaling (Griffin, 1954).
Although enormous problems are associated with scaling, salt water, nevertheless,
finds many applications in the petroleum industry. Oil production in many oil fields
depends on hydrodynamic drive or water compressibility of salt water, which is
responsible for the internal energy of the reservoir. The oil industry recognised the
benefits of water in improved oil recovery methods, such as simple water flooding and
chemical flooding processes. For many of these processes, it is generally not possible to
find appropriate water, which is compatible with the existing formation water.
These incompatibilities between waters can result in substantial scaling on the surface
facilities and reservoir formation matrices. Deposition of mineral scales on the surface of
formation matrices can cause severe formation damage and plugging. The maximum
amount of scale, which may be formed depends on the thermodynamics of mixed waters,
while the extent of deposition depends on the kinetics of the deposition process. Mineral
scales, which consist of various forms of calcium, barium and strontium sulphates are
generally very hard and difficult to be removed. Many cases of oil well scaling have been
reported in the literature; a comprehensive review of these cases is given elsewhere
(Moghadasi et al., 2006). To remove the scales, regular acid stimulation of the formation
adjacent to the production wells and acid washing of surface facilities are required.
The calcium sulphate scale is generally removed with dilute hydrochloride or sulphuric
acid solutions, which cause severe corrosion of facilities and tubing, thus necessitating
frequent maintenance and replacement.
A thorough understanding of the mechanisms of scale formation resulting from
precipitation of inorganic salts such as calcium sulphate in porous media is required to
improve reservoir management and production operation by minimising scale
deposition problems. Furthermore, this information and the knowledge of the
physical and chemical nature of the deposits are needed to identify appropriate
cleaning methods and schedules. The deposit layer formed during the scaling process is
obviously the starting point for any in-depth study of cleaning mechanisms.
However, there is a lack of experimental evidence and fundamental understanding of
scaling and cleaning mechanisms for mineral deposits. This statement is particularly
correct with respect to the availability of data for scaling and cleaning in porous media.
It is generally believed that the best understanding of scaling and cleaning mechanisms
can be obtained by laboratory research on well-defined systems where various operating
parameters such as injection rate, temperature and concentration can be varied to identify
their effects. The aim of the present investigation is, therefore, to study the mechanisms
of scaling and cleaning of calcium sulphate dihydrate formed in porous media
by measuring pressure drop and permeability reduction during scaling and cleaning
periods. The operating parameters such as injection rate, temperature and supersaturation
concentration are varied. After clarification of the mechanisms of scaling and
cleaning, predictive models for calcium sulphate scaling and cleaning with dilute acid
solutions have been developed. The presented models can be easily adapted for other
mineral scales.
84 M. Jamialahmadi and H. Müller-Steinhagen

2 Experimental conditions

Figure 1 shows the schematic layout of the experimental set-up. The test rig was
designed to operate in a temperature range of 50–90°C, similar to that in oil fields.
The basic components of the test rig are test section, circulation pump, tanks and flow
meters. The solution is pumped from the reservoir tank to the test section using a
peristaltic pump. Adjusting the revolutions of the pump controls the flow rate of the
liquid phase. The pump speed ranges from 0.5 to 55 rpm corresponding to flow rates
between 0.2 and 3 cm3/s. These flow rates were intended to reproduce brines flowing
through the formation matrix immediately adjacent to the producing wells, where scale is
expected to be most damaging. The maximum design pressure of the pump is 4 bar,
which is achieved by utilising tubing with an internal diameter of 1.6 mm and a wall
thickness of 1.6 mm. The tubing is made from Marprene II, a material that is resistant to
water and crude oil. The reservoir tanks are made of stainless steel, which can be heated
to 90°C using four band heaters. Each band heater is 40 mm wide and has a power rating
of 500 W at 240 volts. The band heaters are heating against a small cooling coil, which is
placed inside the tanks, to allow a better control of the liquid temperature in the tanks.
Each cooling coil is made from stainless steel tubing with an outside diameter of 6.4 mm,
and a length of 38.1 cm. Rotameters are installed to control the cooling water flow rates.
The temperature of the liquid inside the tanks is determined using thermocouples, which
are connected to a controller, which in turn will control the power output of the band
heaters. This will maintain the temperature at a desired set point. A stirrer is also placed
in each tank to agitate the liquid to provide a uniform temperature and concentration.
The stirrer can rotate between 0 and 2500 rpm.

Figure 1 Porous medium test apparatus


Mechanisms of scale deposition and scale removal in porous media 85

Figure 2 shows the test section, which is made from a stainless steel pipe with 32 mm
diameter and 580 mm length. The pipe can be filled with glass beads of different
diameters, hence forming porous media with different system geometries. Particles
with an average diameter of 0.480 mm are used in this investigation. These particles
provide a bed with porosity and permeability of 0.363 and 1.53 × 10−10 m2, respectively.
Six pressure transducers are installed at equal distances along the bed, which can
measure pressures up to 5 bar with high accuracy. At the inlet and outlet of the test
section, two knit-meshes are kept in place by flanges to contain the particles and prevent
them from leaving the bed. All tubing and fittings are made of stainless steel.
The temperature of the porous medium is controlled by a Thermocoax heating wire,
which is placed in a spiral groove around the test section and embedded by high
temperature soldering tin to ensure good contact with the test section wall. Longitudinal
grooves accommodate thermocouples for measuring the wall temperatures. The bed
expands above and below the heated section to ensure uniform distribution of the flow.
Bulk temperature is measured using six thermocouples, which are inserted along the
length of the bed. Another eight thermocouples are used to measure the wall temperature
at two different positions along the bed. All measurements are fed into a data acquisition
system, which is connected to a desktop computer.

Figure 2 Schematic diagram of porous medium

In order to start an experiment, clean, dried and preweighted particles of a given size are
poured into the test section and placed into a sonic vibrator for one hour. Then, the bed is
levelled, its height recorded and the test section installed back to its position in the test
loop. The pore volume of the dry porous medium was filled with liquid supplied from a
burette connected to the inlet of the bed. Taking the necessary corrections into account,
the porosity of the medium was then calculated as the ratio of the amount used to fill the
pore volume divided by the total volume of the bed. The same procedure was repeated
several times for each medium and the mean value was taken to represent the porosity of
86 M. Jamialahmadi and H. Müller-Steinhagen

the medium. Calcium sulphate was chosen as the dissolved salt because its scale is very
common in the oil industry. The solubility of calcium sulphate in distilled water is shown
in Figure 3 as a function of temperature. At temperatures higher than 40°C, the solubility
of calcium sulphate decreases with increasing temperature. Under normal operating
conditions, if there are enough calcium and sulphate ions in the water, nucleation and
crystallisation and, therefore, scaling of calcium sulphate is expected to take place on the
grain surfaces of the porous medium.

Figure 3 Solubility curves of calcium sulphate in water

In the scaling experiments, aqueous solutions of Na2SO4 and Ca(NO3)2 4H2O were used
to provide these ions. The criteria for selecting these salts were based on solubility of the
salts and valence of the respective ions. Solutions were prepared for each run by
dissolving the salts in distilled water and allowing the solutions to equilibrate for 12 hr.
Each test liquid was hence obtained by mixing two solutions, one rich in calcium ions
and the other rich in sulphate ions, which were kept separate until entering the porous
medium. All salts used in this work were Merck reagent grade. The range of salt
concentrations used in this investigation is given in Table 1.

Table 1 Range of operating parameters

Salt flow rate 0.15–3 cm3/s


Temperature 20–90°C
System pressure 1–3 bar
Na2SO4 concentration 0–3 kg/m3
3
Ca(NO3)2⋅4H2O concentration 0–4.5 kg/m
Degree of supersaturation 0–0.65 kg/m3
Sulphuric acid concentration 5 wt%

After the calcium nitrate and sodium sulphate solutions of predetermined concentrations
have been filled into the supply tanks, distilled water was fed through the bed until the
system reached a steady state at the desired temperature and flow velocity. Once the
system had reached the desired condition, the data acquisition system was switched on to
record fluid flow rate, temperatures and pressures. At this time, the supply pumps were
Mechanisms of scale deposition and scale removal in porous media 87

switched to the scale forming solutions. This procedure was repeated for different
operating variables, which are also given in Table 1.
Immediately after the end of each scaling run, the supply pumps were switched to the
scale removal reagent solution and flooding continued for several hours. This was to
study the removal rate from the scaled porous medium and to remove all salt forming
ions from the system before the subsequent experiments. The test section was then
removed, cleaned and dried to be ready for the next experiment. The experimental runs
were performed in an arbitrary sequence and some experiments were repeated to check
the reproducibility of the results, which proved to be good.

3 Experimental results

3.1 Clean bed experiments


Prior to carrying out any tests with salt solutions, it is necessary to characterise the
porous medium in order to have adequate information regarding the mechanisms and
flow regimes in the clean bed for comparison with the scaled bed. When salt-containing
solution flows through a porous medium, the developed pressure drop along the bed in
the direction of the flow is a function of system geometry, bed porosity and physical
properties of the solution. The flow rate and pressure drop profiles should be known
before considering the mechanisms of scaling in the porous medium. The operating
conditions can result in four distinct flow regimes namely Darcy or creeping flow,
inertial flow, unsteady laminar flow and chaotic flow. The first two regimes are generally
encountered in oil reservoirs; the Darcian flow regime occurs away from the production
wells and the non-Darcian regime in the vicinity of the production wells. The classical
steady-state equation of motion through a porous medium can be expressed as Nield and
Bejan (1999):
−∇P = a0 µ v + a1 ρv × ∇v (1)

Using scale analysis technique, the x-directed component of Equation (1) becomes:
dP
− = a0 µu + a1 ρu 2 (2)
dx
This equation is known as Dupuit and Forchheimer correlation. The two terms on the
right hand side of Equation (2) can be recognised as viscous and inertial contributions.
The parameters a0 and a1 are constants and should be determined experimentally. The
Kozeny-Carman (1927) correlation, which is one form of the Dupuit and Forchheimer
equation, is still the most widely used correlation for the prediction of pressure in
homogeneous unconsolidated porous media:
∆P 150(1 − φ )2 1.75 (1 − φ ) 2
= µu + ρu (3)
L φ dp
3 2
φ 3 dp

At low Reynolds numbers, the inertial term is negligible; this is under creeping flow
conditions where Darcy’s law holds and Equation (2) simplifies to:
dP µ
− = u (4)
dx k
88 M. Jamialahmadi and H. Müller-Steinhagen

A general expression for permeability of porous media can be obtained for the entire
range of Darcian and non-Darcian flow regimes by reducing Equations (3) and (4):

φ 3 dp2
k= (5)
6(1 − φ )2 ( 25 + 0.292 Re m )

In the Darcian flow regime, Rem is small and Equation (5) reduces to:

φ 3 dp2
k= (6)
150(1 − φ )2

Equations (5) and (6) show that the permeability in the Darcian flow regime is
independent of Reynolds number, while for non-Darcian flow regime, it is a function of
fluid Reynolds number. Several other empirical correlations have also been developed to
describe the flow through porous media. These correlations are summarised in
Table 2. To put the accuracy of these correlations into perspective, their predictions are
compared with the experimental data in Figure 4. While all correlations predict an
increase in pressure drop with increasing fluid velocity, the variation between values
from different correlations is quite considerable. The best agreement between measured
and calculated values is obtained from the correlations of Kozeny-Carman (1927), Ergun
(1952), Joseph et al. (1982) and Jamialahmadi et al. (2005), for both Darcian and
non-Darcian flow regimes. As expected, Darcy’s model predicts the experimental data
only at low fluid Reynolds numbers where its prediction coincides with the
Kozeny-Carman correlation.

Table 2 Correlations suggested for the prediction of pressure drop in porous media

Reference Correlation

Darcy (1856) ∆P µ
= u
L k

Single-phase, Darcian flow regime


Blake (1922) ∆P k µ a G 2 SP
= × ×
L g gc ρ φ 3

Single-phase, Darcian and non-Darcian flow regimes


Kozeny-Carman (1927) ∆P 150(1 − φ )2 1.75(1 − φ ) 2
= u+ ρu
L ds2φ 3 φ 3ds

Single-phase Darcian and non-Darcian flow regimes

Leva (1947) 1.9


⎛ µ 2λ 1.1 ⎞
∆P k (1 − φ ) ⎛ dpG ⎞
= ⎜ ⎟ ⎜ ⎟
L gcφ 3 ⎝ µ ⎠ ⎜ ρ dp3 ⎟
⎝ ⎠

Single-phase, Darcian and non-Darcian flow regimes


Mechanisms of scale deposition and scale removal in porous media 89

Table 2 Correlations suggested for the prediction of pressure drop in porous


media (continued)

Reference Correlation
Ergun (1952) ∆P 150(1 − φ )2 1.75(1 − φ ) 2
= u+ ρu
L ds2φ 3 φ 3 dP
Single-phase, Darcian and non-Darcian flow regimes
Joseph et al. (1982) ∆P µ 0.55 2
= u+ ρu
L k k
Single-phase, Darcian and non-Darcian flow regimes
Jamialahmadi et al. (2005) ∆P f 6(1 − φ ) 2 f 25
= ρ u where = + 0.292
L 2 dPφ 3
2 Re

Single-phase, Darcian and non-Darcian flow regimes

80
Correlations : Experiment
(1) : Blake (1922)
(1)
Pressure drop per unit bed height (kPa/m)

(2) : Leva (1947) (2)


(3): : Joseph et al. (1982)
(4) : Kozeny – Carman (1927)
60
(5) : Ergun (1952)
(6) : Jamialahmadi et al. (2005)
(7) : Darcy (1856)

(3) (4), (5), (6)


40

(7)

20

dC = 3.2 cm
dS = 0.48 mm
0
0 2 4 6 8 10 12
Fluid Reynolds number, Rem
Figure 4 Prediction of single phase pressure drop from various correlations

4 Salt-containing solutions

The mechanism of calcium sulphate scaling in porous media is complicated and


depends on various operating parameters, particularly on injection flow rate, temperature
and degree of supersaturation with respect to the concentration of the scale forming
species.
90 M. Jamialahmadi and H. Müller-Steinhagen

4.1 Effect of injection rate


To identify the controlling mechanism of scaling, the pressure drop and its corresponding
permeability reduction of the porous medium are measured over a wide range of
injection rates under constant temperature and calcium sulphate super saturation.
The flow rates varied from 0.15 to 3 cm3/s, a range of flow rates which generally occurs
in the formation close to the production wells. Scaling due to brine flowing in these
regions is expected to be high. During each test, pressure drop across the test section was
recorded as a function of time. Typical results of these measurements are summarised in
Figure 5(a) as a function of injection period for different injection rates under constant
temperature and salt supersaturation concentration. The pressure drop increases during
the experiments, indicating the scale formation on the surface of the grains of the porous
medium. The pressure drop curves show that the variation is more pronounced at high
flow rates and that the rate of scaling increases with increasing flow rate. To estimate the
formation damage, the permeability reduction corresponding to the measured pressure
drop is calculated from Darcy’s law and the results are presented in Figure 5(b).
The permeability ratio curves show an almost linear decrease in permeability with time.
This is especially true for the initial time of scaling, for all flow rates of this
investigation. Helalizadeh (2002) has shown that for crystallisation scaling, the attractive
forces between crystalline scale and the surface are much stronger than the shear forces
due to the fluid. This indicates that in crystallisation scaling, no significant erosion
processes are to be expected, even at higher injection rates. Figure 6 shows the variation
of initial rate of permeability reduction with injection rate for constant temperature and
supersaturation concentration. The results show that as injection rate increases the scaling
rate and, therefore, the rate of reduction of permeability increases.

Figure 5 (a) Pressure drop as a function of time showing the effect of injection rate
Mechanisms of scale deposition and scale removal in porous media 91

(b)
Figure 5 (b) Variation of permeability ratio with time at various injection rates at constant
bulk concentration and temperature (continued)

Figure 6 Region of mass transfer limited and reaction limited scaling

4.2 Effect of temperature


The role of temperature on the solubility of scale and the morphology of crystal growth
is very important. The lattice structure of the calcium sulphate, which is formed on the
surface of grains of the porous medium at different temperatures varies widely.
Three distinct lattice structures have been reported for calcium sulphate scales, namely
92 M. Jamialahmadi and H. Müller-Steinhagen

gypsum or dihydrate, bassanite or hemihydrate and anhydrite. These crystals differ


chemically by the amount of water contained in their structures. Anhydrite crystals are
shorter and have the closest packing, highest density and most stable arrangement of any
calcium sulphate deposit. Furthermore, the anhydrite deposit is harder, more adherent
and, therefore, more difficult to be removed from the substrate surface than dihydrate
and hemihydrate deposits. The phase transitions of calcium sulphate are shown in
Figure 7. During the present investigation, the operating temperature was maintained
between 50°C and 80°C; therefore, calcium sulphate dihydrate predominate the scale
formed on the surface of grains in the porous medium. The appearance of calcium
sulphate dihydrate crystals formed at different temperatures and flow rates varies, as can
be seen from the photographs in Figure 8. At low temperature and flow rate, large
needle-like crystals are observed. With increasing temperature and flow rate, the
crystals become small, thin and plate-shaped. Deposits formed at high temperature and
flow rate seem to be harder and more adherent than deposits formed at low temperature
and flow rate.

Figure 7 Phase transition of calcium sulphate deposit with temperature

Figure 8 Scale observed for two different injection rates and temperatures

Figure 9(a) and (b) shows typical variations of pressure drop and their corresponding
permeability reduction with time, for different temperatures at constant injection rate and
salt concentration. The rates of increase of pressure drop and of permeability ratio
decline increase strongly with operating temperature. Figure 10 shows the rate of
reduction of permeability with temperature at two different flow rates under constant
degree of supersaturation. The results show that the rate of reduction of permeability
increases with increasing temperature and that this trend is more pronounced at higher
flow rate. The main effects of temperature on mass transfer of a species towards the
grains surfaces are through the diffusion coefficient of the scaling ions, the saturation
concentration and the reaction rate constant of the scale forming reaction. Since mass
transfer coefficients are a linear function of temperature, the scaling rate should increase
linearly with temperature. This trend was observed at low flow rates. If the injection rate
Mechanisms of scale deposition and scale removal in porous media 93

is increased, the effect of temperature should become more pronounced, approaching the
Arrhenius relationship for reaction controlled processes. The present results show that
while the effect of temperature becomes stronger with increasing liquid flow rates, it
almost never reaches the condition where scaling is entirely surface reaction controlled.
Therefore, to develop a predictive model for scaling and permeability reduction, the
combined effects of mass transfer and reaction should be taken into consideration.

Figure 9 (a) Pressure drop as a function of time showing the effect of temperature and
(b) Variation of permeability ratio with time at various temperatures under
constant bulk concentration and injection rate
94 M. Jamialahmadi and H. Müller-Steinhagen

Figure 10 Variation of permeability ratio reduction with temperature for different injection
rates at constant supersaturation concentration

4.3 Effect of concentration


The primary cause of scale formation is the supersaturation of the solution with
respect to the dissolved salts. When the concentration product of calcium and
sulphate ions exceeds the saturation value, calcium sulphate precipitates and forms
scale. As long as shear-related removal can be ignored, the rate of scaling can be
expressed by:

( )
n
m s = kr CAi − CA* (7)

where CA* is the saturation concentration of calcium sulphate under the grain surface
conditions. For mass transfer and reaction controlled deposition, n is equal to one and
two, respectively. Equation (7) shows that regardless of the mechanism of scaling, the
effect of concentration is important and that it should be more pronounced for reaction
controlled conditions. Figure 11(a) and (b) shows the variation of pressure drops and
their corresponding permeability ratio reduction with time at constant injection rate and
temperature for different concentrations of calcium sulphate. The results illustrate that by
increasing the concentration of calcium sulphate ions, the degree of supersaturation of
the solution increases and, therefore, the formation of scale and the reduction of
permeability ratio are accelerated.
Mechanisms of scale deposition and scale removal in porous media 95

800
CAb (kg/m3) Experiment Q (cm3/s) T (oC) Equation (3)
2.8 1.333 80
2.5 1.333 80
Pressure drop, per unit height (Pa/m)

2.2 1.333 80
600 80
2.0 1.333

400

200

0
0 10 20 30 40 50
Time (hr)
(a)

1.0

0.8
Permeability ratio, Rk

0.6

0.4 CAb (kg/m ) Experiment


3 Q (cm3/s) T (oC) Presented model
2.0 1.333 80
2.2 1.333 80
2.5 1.333 80
2.8 1.333 80
0.2
0 10 20 30 40 50
Time (hr)
(b)

Figure 11 (a) Pressure drop as a function of time showing the effect of concentration and
(b) Variation of permeability ratio with time at various concentrations under
constant temperature and injection rate

5 Mechanism of scaling

Scaling of calcium sulphate dihydrate on a grain surface of the porous medium can be
expressed by the following reaction:
Ca 2 + + SO24− + 2H 2 O → CaSO 4 2H 2O ↓ (8)
96 M. Jamialahmadi and H. Müller-Steinhagen

Almost all investigators have shown experimentally that the above reaction is an
elementary second order reaction in plain tubes (i.e. Bohnet, 1987) and porous media
(i.e. Moghadasi, 2002). The rate of formation of calcium sulphate scale per unit surface
area of the grains can be presented as:
V dC
(
= kr ⎡⎣Ca 2 + ⎤⎦ ⎡⎣SO24 − ⎤⎦ = k0 e − ( E / RT ) CAi − CA* )
2
m s = (9)
A dt
The deposition of calcium sulphate on the grain surface takes place in two consecutive
steps: in the first step, calcium and sulphate ions are transported to the surface by
diffusion processes through the boundary layer, which is formed between the grain
surface and the salt solution in the pores:
m s = β ( CAb − CAi ) (10)

CAi is the concentration of calcium sulphate on the grain surfaces, which is somewhat
between the bulk concentration, CAb and the saturation concentration, CA* . Once the ions
reach the reaction zone on the grain surface, they react according to Equation (8) to form
crystalline calcium sulphate at the grain surface–fluid interface, which constitutes the
scale. As these two steps occur in series, Equations (9) and (10) can be combined to
eliminate the interface concentration:
⎛ β2 ⎞
( ) ( )
2
m s2 − ⎜ + 2β CAb − CA* ⎟ m s + β 2 CAb − CA* =0 (11)
⎝ kr ⎠

The scaling rate m s can be obtained from the quadratic Equation (11) in terms of mass
transfer coefficient, bulk and saturation concentrations:
⎡ ⎛ β ⎞
2 ⎤
β β
m s = β ⎢
⎢ 2 kr
(
+ CAb − CA − ⎜
*
) ⎟
⎝ 2 kr ⎠
+
kr
CAb − CA* ( ) ⎥⎥ (12)
⎣ ⎦
Equation (12) predicts the scaling rate at any position in the porous medium on the grain
surfaces in terms of mass transfer coefficient, bulk and saturation concentrations of ions.
These variables should be determined before calculating the scaling rate from
Equation (12).

5.1 Mass transfer coefficient


A classical expression for computing mass transfer coefficients for liquid flow through
porous media is the correlation of Geankoplis (1983):
1.09
JD = Re −2 / 3 (13)
φ
JD is the Chilton and Colburn J-factor, which is equal to:
β
JD = Sc2 / 3 (14)
u
Mechanisms of scale deposition and scale removal in porous media 97

This model is valid for Reynolds numbers from 0.0016 to 55 and Schmidt numbers from
165 to 10,690, which covers both Darcian and non-Darcian flow regimes. Combining
Equations (13) and (14) yields:
−2 / 3
1.09u ⎛ udp ⎞ 1.09u
β= = Pe −2 / 3 (15)
φ ⎜⎝ D ⎟⎠ φ

Equations (15) and (12) show that the scaling rate increases as injection rate increases if
the scaling process is mass transfer controlled, while for reaction controlled scaling the
deposition rate is independent of flow velocity as long as the temperature remains
constant.

5.2 Calculation of scale solubility


The solubility of a given salt in water is generally controlled by temperature, pressure,
pH and the concentrations of other substances in water. Once the solubility of a dissolved
salt such as calcium sulphate is exceeded, some of the dissolved salt will precipitate out
to form scale. A variety of solubility prediction models have been recommended for
common oil reservoir scale-forming minerals, which are surveyed and critically
discussed elsewhere (Jamialahmadi and Müller-Steinhagen, 2006). Oddo and Tomson
(1994) developed a relatively simple correlation for calculating the solubility of various
forms of scales. The saturation ratio SR is defined as the ratio of the ions activity product
to the solubility product:
aMe × aAn
SR = (16)
KSP

where KSP is the solubility product of the salt, aMe and aAn are the activity of the cations
and anions of the salt, respectively. The activity of a species is defined as the product of
its activity coefficient and molar concentration:
ai = [ Ci ] γ i (17)

γi is the activity coefficient of an ion, which can be calculated from Skoog (1992):
0.5085Z i2 I
log γ i = − (18)
1 + 0.3281α i I

Substituting of Equation (17) into Equation (16) and rearranging yields:


γ Meγ An
SR = [Me][An] (19)
KSP

Activity coefficients are a function of temperature, pressure, ionic strength and effective
diameter of the hydrated ion αi. Unfortunately, considerable uncertainty exists regarding
the magnitude of αi in Equation (18). To eliminate this error, Odd and Tomson (1994)
included the activity coefficients in the solubility product KSP. The advantage of this
inclusion is that the activities of the respective cations and anions of interest need not be
determined directly, but are included implicitly in the constants, which can be obtained
from the available experimental data. Therefore, Equation (19) becomes:
98 M. Jamialahmadi and H. Müller-Steinhagen

[Me][An] γ γ
SR = ′ = Me An
where K SP (20)

K SP KSP

Scaling cannot occur unless the solubility of the salt in question is exceeded. The scaling
index defined by Equation (21) uses the chemistry of an aqueous solution to indicate
whether the solution has the potential for scaling:
SI = log(SR) (21)

The general saturation index for the scale deposit can be obtained by substituting
Equation (20) into Equation (21):
SI = log (SR) = log {[Me][An]} − log K SP
′ (22)

Writing Equation (22) in terms of pKSP gives:



SI = log{[ Me][ An ]} + pKSP where pK SP = − log KSP (23)

The parameter pKSP is a function of temperature, pressure and ionic strength.


This functionality is expressed as:
pK SP = b0 + b1T + b2T 2 + b3 P + b4 I 0.5 + b5 I + b6 I 0.5T (24)

Substituting Equation (24) into Equation (23) reads:

{ }
SICaSO4 = log ⎡⎣ Ca 2 + ⎤⎦ ⎡⎣SO 24 − ⎤⎦ + b0 + b1 T + b2T 2 + b3 P + b4 I 0.5 + b5 I + b6 I 0.5 T (25)

The parameters of Equation (25) for calcium sulphate scales are summarised in Table 3.
Coefficients for other salt scales can be found elsewhere (Jamialahmadi and
Müller-Steinhagen, 2006; Oddo and Tomson, 1994).

Table 3 The parameters of Equation (25) for prediction of CaSO4 saturation index

Salt scale b0 b1 b2 b3 b4 b5 b6
CaSO4.2H2O 3.47 1.8 × 10−3 2.5 × 10−6 −5.9 × 10−5 −1.13 0.37 −2.0×10−3
1
CaSO4. H2O 4.04 −1.9 × 10−3 11.9 × 10−6 −6.9 × 10−5 −1.66 0.49 −0.66×10−3
2
CaSO4 2.52 9.98 × 10−3 −0.97 × 10−6 −3.07 × 10−5 −1.09 0.5 −3.3×10−3

A negative saturation index indicates a non-scaling condition, a zero index indicates an


equilibrium condition and a positive index indicates that scale may form in the
formation, well tubing and surface facilities.

6 Modelling of scale deposition

While the understanding of the mechanisms of scaling in porous media has improved,
sound modelling of the reduction of permeability of the porous media with time due to
scale formation is not possible yet. The scaling rate is a function of time and location in
the porous medium. It is often difficult and time consuming to determine the rate of
Mechanisms of scale deposition and scale removal in porous media 99

scaling at a given position in the system at a particular time, but the average mass of salt
deposited in the medium can be calculated. The total mass of scale deposited on the
surface of the grains between time zero and time t can be expressed as:
ms = m s × Sp × t (26)

where Sp is total external surface area of the grains, which is equal to:
S p = N p π d p2 (27)

The total number of grains Np can be related to the bed porosity by the following
equation:

π dp3 6 (1 − φ0 ) Vb
VP = (1 − φ0 ) Vb = N P or N P = (28)
6 π dp3

Replacing Equations (27) and (28) into Equation (26) gives:


6 (1 − φ0 ) Vb
ms = m s t (29)
dp

The volume of the scale is the ratio of the mass of scale given by Equation (29) divided
by the density of scale:
6 (1 − φ0 ) Vb
Vs = m s t (30)
ρs dp

For the density of the CaSO4, a typical value of 1990 kg/m3 may be taken
(Müller-Steinhagen and Jamialahmadi, 2006). Finally, the fraction of the bulk volume of
the porous medium that is occupied by scale is equal to:

Vs 6 (1 − φ0 )
φs = = m st (31)
Vb ρs d p

Therefore, the porosity ratio at time t can be calculated from:

φt φ 6 (1 − φ0 )
= 1− s = 1− m t (32)
φ0 φ0 ρs d pφ0 s

Substituting for m s from Equation (12) yields:

6β (1 − φ0 ) ⎡ β ⎤
2
φt ⎛ β ⎞ β
= 1− ⎢ + ( CAb − CA* ) − ⎜ + ( C − C A) t
* ⎥
(33)
φ0 ρs d pφ0 ⎢ 2 kr ⎝ 2 kr ⎟⎠ kr Ab ⎥
⎣ ⎦
Permeability ratio can be related to porosity ratio by writing Equation (6) for clean and
scaled porous media and then dividing the two equations:
3 2
k ⎛ φt ⎞ ⎛ 1 − φ0 ⎞
= (34)
ki ⎜⎝ φi ⎟⎠ ⎜⎝ 1 − φt ⎟⎠
100 M. Jamialahmadi and H. Müller-Steinhagen

Equation (33) in conjunction with Equation (34) may be used to predict the porosity
and permeability of the porous medium as a function of scaling period. Predictions
of these equations have also been included in Figures 5(b), 9(b) and 11(b).
The calculated trends are in good agreement with the experimental data. The absolute
mean average error between the proposed models and the experimental data is
about 11%, which reflects the acceptable agreement with the experimental data
and the suitability for the prediction of formation damage due to scaling of
inorganic salts.

7 Scale removal

The process of scale removal by acid, also referred to as acid stimulation, involves
the injection of dilute sulphuric or hydrochloric acid solution into the subsurface
formation. Almost no satisfactory chemical method has been invented to-date to
completely eliminate scale formation in porous media. Therefore, sooner or later,
it will be inevitably required to remove the formed scale in order to sustain the
production of oil from the reservoir. Scale removal to increase the productivity of oil
wells is a common process in the petroleum industry, even though steps have been
taken to minimise the potential of scaling problems at design and during subsequent
operation. A wide range of acids and other chemicals are used for scale removal and
cleaning in the oil industry. Theoretically, calcium carbonate scale can be controlled
by acid injection or by addition of acid-forming salts such as ferric chloride (York
and Schorle, 1966). Among the various types of acids that have been proposed and
used for scale removal, sulphuric acid may be suitable because of its cost and
concentration. Other strong acids, such as hydrochloric have been used to control pH, but
these acids present more problems in cost, handling and corrosion than sulphuric acid.
The acid interacts with the deposit within the porous medium and dissolves the
scale, thereby increasing the porosity and permeability of the pores for flow of oil.
The solubility of calcium sulphate in 5 wt.% sulphuric acid is given in Figure 12.
The solubility of calcium sulphate scale in acids is a strong function of the pH of the
solution (Helalizadeh, 2002). Calcium sulphate deposits are almost insoluble in
inorganic acids as long as the pH of the solution is in the range of 7–3, but its solubility
increases as the pH value is reduced further. Sulphuric acid with concentration of 5 wt.%
has a pH-value close to pH = 0 and is hence suitable as a cleaning solution. Furthermore,
Figure 12 shows that the solubility of calcium sulphate in sulphuric acid increases
with increasing temperature. Typical results of scale-removal experiments are shown in
Figure 13. The results show that the pressure drop across the system is decreased
gradually as the dissolution process proceeds, confirming that scale is removed and,
therefore, the permeability of the porous medium is improved. The permeability ratio
increases towards an asymptotic value of 1 for the clean system. Similarly to the
problem of scaling, effective removal of scales by acid also depends on injection rate
and temperature of acid solution. High velocity and elevated temperature of the scale
removal operation will assist the removal process by decreasing the boundary layer
thickness between bulk acid flow and scale surface, and by increasing the solubility of
the scale in acid.
Mechanisms of scale deposition and scale removal in porous media 101

Figure 12 Solubility of calcium sulphate in dilute sulphate acid solution as a function of


temperature

Figure 13 Typical variation of pressure drop and permeability ratio during the scale removal
process
102 M. Jamialahmadi and H. Müller-Steinhagen

7.1 Effect of injection rate and temperature of acid solution


The effect of injection rate of acid solution on the removal of calcium sulphate scale is
illustrated in Figure 14. Since the scaling experiments were stopped once the
permeability ratio had decreased to a preselected value of about 0.46, the initial
permeability ratio for scale removal tests was almost constant. The results show that as
the injection rate of the acid solution is increased, the removal rate of scale is increased
and, therefore, the cleaning period is decreased. Mass transfer rates of the acid from the
bulk of the solution in the pores to the solution/scale interface and of dissolved deposit
from the scale surface to the bulk of the solution are both increasing with decreasing
boundary layer thickness, hence reducing the remaining mass of scale and the removal
time requirement. The effect of temperature on the dissolution of scale in the acid
solution is shown in Figure 15 for constant injection rate. There are two physical effects
related to the temperature, which influence the dissolution of scale in the sulphuric acid
solution:
1 as Figure 12 shows, by increasing the temperature, solubility of calcium
sulphate scale in dilute sulphuric acid solution is increased and, therefore, the
removal period is decreased
2 since the mass transfer coefficient is a linear function of temperature, the
removal rate of scale should increase with temperature.

Figure 14 Effect of injection rate on scale removal process at constant temperature


Mechanisms of scale deposition and scale removal in porous media 103

Figure 15 Effect of temperature on scale removal process at constant injection rate

8 Modelling of scale removal process

The effectiveness of a chemical to remove scale from the surface of the grains of the
porous medium depends on its ability to dissolve the scale over a reasonable time
without damaging the other facilities. The dissolution process of calcium sulphate scale
in dilute sulphuric acid may be expressed as:
CaSO 4 × 2H 2 O( S ) → CaSO 4( aq) + 2H 2 O (35)

The rate at which the calcium sulphate scale dissolves in the acid solution can be
obtained from the integral form of the one-dimensional equation of continuity in porous
media (Treybal, 1980):

m r = rA ln
( ( ))
Deff Cav rA 1 + Deff / DK,eff Cav − CAb
(36)
zf ( ( ))
rA 1 + Deff / DK,eff Cav − CAi

According to the film theory of Whiteman (1923), the interfacial concentration of


calcium sulphate CAi is equal to its solubility CA* at the given conditions. The parameter
rA in Equation (36) is equal to:
NA
rA = (37)
NA + NB

The stoichiometry of the dissolution reaction fixes the relationship NB = 2NA, and
Equation (37) gives rA = 1/3. It has been shown (Scott and Dullin, 1962) that molecular
104 M. Jamialahmadi and H. Müller-Steinhagen

diffusion prevails for liquid flow through porous media (i.e. Dk,eff >> Deff). The effective
diffusivity Deff is smaller than the normal diffusivity D, to which it is related through bed
tortuosity (Cussler, 1985):
D
Deff = (38)
τ
Tortuosities usually range between 2 and 6 (Cussler, 1985), even though values as high
as 14 have been reported. For the present investigation, a value of 4.5 was used in the
following calculations. Substituting Equation (38), the value of rA = 1/3 and the mass
transfer coefficient (i.e. β = D/zf) into Equation (36) yields:
1 βCav Cav − 3CAb
m r = ln (39)
3 τ Cav − 3CA*

Equation (39) shows that the rate of removal of scale during acid cleaning increases as
temperature, saturation concentration and injection rate increases, which is in complete
agreement with the experimental observations. Hence, the rate of improvement of the
porosity ratio at any time is equal to:

φ φi 6 (1 − φi )
= + m t (40)
φ0 φ0 ρs d pφ0 r

where φi/φ0 is the porosity ratio at the start of scale removal process. Predictions of
Equation (40) in conjunction with Equations (39) and (34) have also been included in
Figures 13–15 for the investigated injection rate of the acid solution. The calculated
trends are in good agreement with the experimental results. The absolute mean average
error between the suggested model and the experimental data is about 14%. Considering
the complexity of scale removal processes, this can be considered as a good match with
the experimental data. Nevertheless, it should be considered that the above model only
applies for scale removal processes, which do not experience significant shear-related
removal of scale.

9 Conclusions

Scaling and scale removal are two common problems that are categorised among the
most expensive processes in petroleum industry. Calcium sulphate dihydrate scaling
increases the pressure drop and decreases the permeability of the porous media
considerably. Injection rate, temperature and concentration of salts have the main
influence on the scaling process. The scaling rate increases as concentration and
temperature increase. Depending on the injection rate, the scaling rate is controlled by
different mechanisms. At low injection rates, the mass transfer of reactants to the grain
surfaces has a significant effect on the scaling process, while at higher injection rates, the
scaling is influenced mostly by the chemical reaction on the grain surfaces.
The effects of surface temperature, sulphuric acid concentration and injection rate on
the rate of scale removal are tested. It has been shown that the scale removal process is
controlled by mass transfer from the cleaning acid to the scale surface and of the reaction
product away from the grain surfaces. Finally, the experimental results are used to
develop mechanistic models for the prediction of the rate of permeability reduction due
Mechanisms of scale deposition and scale removal in porous media 105

to scaling and for the removal rate of scale by dilute sulphuric acid solution. Considering
the complexity of the scale formation and removal processes, the quantitative agreement
between measured and predicted values is very good.

Nomenclature

a activity, kg m–3
a0 and a1 constants
b0 – b6 constants
C concentration, kg m−3
dc bed diameter, cm
ds particle diameter, mm
D normal diffusivity, m2 s−1
Deff effective diffusivity, m2 s−1
Dk,eff Knudsen effective diffusivity, m2/s
E activation energy, J mol–1
f friction factor
g gravitational acceleration, m/s2
G mass flux, kg/m2s
I ionic strength, kmol m–3
JD Chilton and Colburn J−factor
k permeability, m2
ki initial bed permeability, m2
k0 Arrhenius rate constant, m4 kg–1 s–1
kr reaction rate constant, m4 kg–1 s–1
Ksp solubility product
L bed height, m
ms mass of scale, kg
m r rate of removal of scale, kg m–2 s–1

m s rate of scaling, kgm–2 s–1

n order of scaling reaction


N molar flow of species, mol m–2 s–1
Np total number of grains in the bed
106 M. Jamialahmadi and H. Müller-Steinhagen

P pressure, kPa
pKsp defined by Equation (24)
Pe Peclet number
rA defined by Equation (37)
R universal gas constant, J mol–1 K–1
Re Reynolds number
Rem modified Reynolds number = ρud p / (1 − φ) µ

Sc Schmidt number
Sp grain surface area, m2
SI Saturation Index
SR Saturation Ratio
t time, s
T temperature, K
u velocity in x−direction, m s–1
v velocity vector, m s−1
V volume, m3
Vb bulk volume of the bed, m3
Vp grain volume, m3
Vs scale volume, m3
x coordinate
zf mass transfer film thickness, m

Greek symbols
αi effective diameter of the hydrated ions, m
β mass transfer coefficient, m s–1
φ porosity
φi initial bed porosity
γ activity coefficient
λ thermal conductivity, W m–1K–1
µ liquid viscosity, kg m–1 s–1
ρ liquid density, kg m–3
ρs scale density, kg m–3
τ tortuosity, defined by Equation (38)
Mechanisms of scale deposition and scale removal in porous media 107

Subscripts and superscript


av average
A solute
An anion
B bulk
Me metal ion
i interface
s scaling
* saturation

References

Blake, F.V. (1922) ‘The resistance of packing to fluid flow’, Transactions on American Institute of
Chemical Engineers, Vol. 14, pp.415–421.
Bohnet, M. (1987) ‘Fouling of heat transfer surfaces’, Chemical Engineering and Technology,
Vol. 10, pp.113–125.
Carman, P.C. (1937) ‘Fluid flow through granular beds’, Industry of Chemical Engineering,
London, Vol. 15, pp.150–162.
Cussler, E.L. (1985) Mass Transfer in Fluid Systems, 1st edition, Cambridge, London: Cambridge
University Press.
Darcy, H. (1856) Les Fontaines Publiques de Ia ville de Dyon, Victor Dalmont.
Ergun, S. (1952) ‘Fluid flow through packed columns’, Chemical Engineering Progress, Vol. 48,
pp.89–105.
Geankoplis, C.J. (1983) Transport Processes, Momentum, Heat and Mass, 2nd edition, USA: Allyn
and Bacon, Inc.
Griffin, A.E. (1954) ‘Water treatment for water flooding’, Producers Monthly, Vol. 18, No. 6,
pp.65–68.
Helalizadeh, A. (2002) ‘Mixed salt crystallization fouling’, PhD Thesis, Department of Chemical
and Process Engineering, University of Surrey, UK.
Jamialahmadi, M. and Müller-Steinhagen, H., 2006 Crystallization fouling, Desware; Encyclopedia
of Desalination and Water Resources, Chapter 12.
Jamialahmadi, M., Müller-Steinhagen, H. and Izadpanah, M.R. (2005) ‘Pressure drop, gas holdup
and heat transfer during single and two-phase flow through porous media’, International
Journal of Heat and Fluid Flow, Vol. 26, pp.156–172.
Joseph, D.D., Nield, D.A. and Papnicolaou, G. (1982) ‘Nonlinear equation governing flow in a
saturated porous medium’, Water Resources Reserch, Vol. 18, pp.1049–1052.
Leva, M. (1947) ‘Pressure drop through porous packed tubes: Part I. A general correlation’,
Chemical Engineering Progress, Vol. 43, pp.713–717.
Moghadasi, J. (2002) ‘Particle movement and scale formation in porous media’, PhD Thesis,
Department of Chemical and Process Engineering, University of Surrey, UK.
Moghadasi, J., Müller-Steinhagen, H., Jamialahmadi, M. and Sharif, A. (2006) ‘Prediction of scale
formation problems in oil reservoirs and production equipment due to injection of
incompatible waters’, Development in Chemical Engineering Mineral Process, Vol. 14,
Nos. 3/4, pp.1–22.
108 M. Jamialahmadi and H. Müller-Steinhagen

Nield, D.A. and Bejan, A. (1999) ‘Convection in Porous Media, 2nd edition, New York:
Springer-Verlag.
Oddo, J.E. and Tomson, M.B. (1994) ‘Algorithms can predict; inhibitors can control scale’,
Oil and Gas Journal, pp.33–37.
Scott, D.S. and Dullien F.A.L., (1962) ‘Diffusion in porous solid’, AIChE Journal, Vol. 8,
pp.113–117.
Skoog, D.A., West, D.M. and Holler, F.J. (1992) Fundamentals of Analytical Chemistry,
6th edition, A Harcourt Brace Jovanovich College Publisher.
Treybal, R.E. (1980) Mass Transfer Operations, 3rd edition. McGraw-Hill book Company.
Whiteman, W.G. (1923) ‘Studies in mass transfer across a gas-liquid interface’, Chemical. and
Material Engineering, Vol. 29 pp.147–154.
York, S. and Schorle, A. (1966) ‘Principle of Desalination’, Chapter 10, K.S. Spiegler.

Potrebbero piacerti anche