You are on page 1of 17

Journal of Petroleum Science and Engineering 43 (2004) 201 217 www.elsevier.

com/locate/petrol

Model study on the kinetics of oil field formation damage due to salt precipitation from injection
J. Moghadasi a,b, H. Muller-Steinhagen c,*, M. Jamialahmadi a, A. Sharif b
a Petroleum Engineering Department, Petroleum University, Ahwaz, Iran Chemical and Process Engineering Department, University of Surrey, Guildford, GU2 7XH, UK c Institute for Thermodynamics and Thermal Engineering, University of Stuttgart, D-70550, Stuttgart, Germany b

Received 16 June 2003; accepted 5 February 2004

Abstract Due to the extensive use of water injection for oil displacement and pressure maintenance in oilfields, many reservoirs experience the problem of scale deposition when injection water starts to break through. In most cases, the scaled-up wells are caused by the formation of sulfate and carbonate scales of calcium and strontium. Due to their relative hardness and low solubility, there are limited processes available for their removal and preventive measures such as the squeeze inhibitor treatment have to be taken. It is therefore important to gain a proper understanding of the kinetics of scale formation and its detrimental effects on formation damage under both inhibited and uninhibited conditions. This paper presents the results of an experimental and theoretical study of permeability reduction of porous media caused by scaling. Two incompatible solutions of calcium and sulfate/carbonate ions were injected into the porous medium, where calcium sulfate or calcium carbonate was generated by chemical reaction. Mechanisms by which a precipitate reduces permeability include solids deposition on the pore walls due to attractive forces between the particles and the surfaces of the pores, individual particles blocking pore throats, and several particles bridging across a pore throat. The characteristics of the precipitate influence the extent of formation damage. Conditions such as large degree of supersaturation, presence of impurities, change in temperature, and rate of mixing control the quantity and morphology of the precipitating crystals. D 2004 Elsevier B.V. All rights reserved.
Keywords: Kinetics; Oil field formation damage; Salt precipitation

1. Introduction Formation damage is a generic terminology referring to the impairment of the permeability of petroleum bearing formations by various adverse processes.
* Corresponding author. E-mail addresses: J.Moghadasi@surrey.ac.uk (J. Moghadasi), hms@itw.uni-stuttgart.de (H. Muller-Steinhagen), A.Sharif@surrey.ac.uk (A. Sharif). 0920-4105/$ - see front matter D 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.petrol.2004.02.014

Formation damage is an undesirable operational and economic problem that can occur during the various phases of oil and gas recovery from subsurface reservoirs including production, drilling, hydraulic fracturing and workovers operations. Formation damage is an expensive headache to the oil and gas industry. The fundamental processes causing damage in petroleum bearing formations are: physico-chemical, chemical, hydrodynamic, thermal and mechanical. Formation damage studies are carried out for

202

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

understanding of these processes via laboratory and field testing, development of mathematical models via the description of fundamental mechanisms and processes. One of the main mechanisms of formation damage is mineral scale formation. The formation of mineral scale associated with the production of hydrocarbons has always been a concern in oilfield operation. Depending on the nature of the scale and on the fluid composition, the deposition can take place within the reservoir which causes formation damage (Khatib, 1994; Krueger, 1986; Leon and Scott, 1987; Lindlof and Stoffer, 1983; Moghadasi et al., 2003) or in the production facilities where blockage can cause severe operational problems. The two main types of scale which are commonly found in the oilfield are carbonate and sulfate scales (Krumine et al., 1985; Lasetr et al., 1968; Nancollas and Campbell, 1969; Nancollas and Liu, 1975; Nancollas et al., 1978; Pang and Sharma, 1994; Singley, 1985; Valone and Skillern, 1982; Vetter, 1972; Vetter and Farone, 1987; Vetter et al., 1982, 1987). While the formation of carbonate scale (Bezmer and Bauer, 1969; Vetter et al., 1982, 1987) is associated with pressure and pH changes of the production fluid, the occurrence of sulfate scale is mainly due to the mixing of incompatible brines, (Leon and Scott, 1987; Lindlof and Stoffer, 1983; Liu and Nancollas, 1971; Moghadasi et al., 2003; Moghadasi et al., 2002; Vetter et al., 1982; Vitthal and Sharma, 1992), i.e. formation water and

injection water. In the oilfield, the universal use of sea water injection as the primary oil recovery mechanism and for pressure maintenance means that problems with scale deposition, (Moghadasi et al., 2002, 2003) mainly with calcium and strontium, are likely to be present at some stage during the production life of the field. A field example is the Iranian offshore Siri field (Moghadasi et al., 2002) in the southern Persian Gulf. The oil in this field comes from a formation called Mishrif, which is common between Iran and the United Arab Emirates. Water injection into the Siri field was started in 1984 with 9100 bbl/day in order to maintain the pressure and to increase the oil recovery. However, the injectivity decreased rapidly; by 1990 the water injection was only 2200 bbl/day, and subsequently, the water injection was stopped. The history of water injection in the Siri field is shown in Fig. 1 (Moghadasi et al., 2002). Table 1 shows the various types of scales that are commonly found in oilfields (Moghadasi et al., 2003) and Table 2 shows the major components of scales found in Iranian and international oilfields (Moghadasi et al., 2003). The deposits are seldom pure calcium sulfate or calcium carbonate, but are usually a mixture of two or more of the inorganic components plus corrosion products, congealed oil, paraffin, silica and other impurities. These materials are trapped in the inorganic lattice and frequently complicate the removal of the deposit. From all the scales, calcium

Fig. 1. Water injection history in the Iranian Siri oil field (Moghadasi et al., 2002).

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217 Table 1 Most common scales in oilfield Moghadaet et al., 2002 Name Calcium carbonate (Calcite) Calcium sulfate: Gypsum (most common) Hemihydrate Anhydrate Barium sulfate Strontium sulfate Iron compounds: Ferrous carbonate Ferrous sulfide Ferrous hydroxide Chemical formula CaCO3 Primary variables Partial pressure of CO2, temperature, total dissolved salts Temperature, total dissolved salts, pressure

203

CaSO4S2H2O

CaSO4S1/2H2O CaSO4 BaSO4 SrSO4 FeCO3 FeS Fe (OH)2

Temperature, pressure Total dissolved salts Corrosion, dissolved gases, pH

sulfate and calcium carbonate have been singled-out for study in this work because: (i) they form a hard scale that cannot be removed easily by chemical treatment; (ii) it is certain that many oil fields are likely to suffer from calcium sulfate and calcium carbonate scale deposition during their lifetime. Calcium sulfate scale, especially gypsum, causes unpredictable problems in many oil fields (Atkinson et al., 1991; Blout and Dickson, 1973; Oddo and Tomson, 1994). This type of scale can cause severe plugging of equipment and producing formations; it creates the necessity of costly remedial work. Thus, to understand the formation of these scale minerals, the kinetics of calcium sulfate scale formation and transformation of one hydrate to another are particularly important in determining the nature of the scale formed under field conditions as a function of both temperature and ionic strength of the solution. 1.1. Previous work The formation of precipitates starts with nucleation processes in which attracted atoms join to form submicron nuclei (Atkinson et al., 1991; Cowan and Weintritt, 1976; Crow, 1985). When impurities are present in the fluid, the energy required to form these

nuclei is smaller than in a pure solution, because the impurities act as nucleation sites. This process, called heterogeneous nucleation, occurs commonly in subsurface reservoirs where fluids contain mineral particles in suspension. In experimental studies, the extent of permeability damage caused by flowing precipitates in cores and sand packs was investigated. Several investigators tested incompatible waters from oilfields in the North Sea (Mitchell et al., 1980; Valone and Skillern, 1982), the Middle East and El-Margon (El-Hattab, 1982) that produced barium, calcium and strontium precipitates. They reported that 15 20-Am-sized crystals blocked the pore throats in glass bead packs and in alumina cores, by size exclusion and bridging mechanisms. Crystals often grew perpendicular to the pore walls, and that some crystal aggregates also had the form of books and rosettes. Todd and Yuan (1991) also conducted laboratory investigations using North Sea reservoir brines that produced barium and strontium scales. Crystals depositing along and growing perpendicular to the pore surfaces caused most of the reduction in core permeability. They observed that doubling the supersaturation ratio of both barium and strontium produced an increase in the quantity of scale formed inside the pores and a change in the morphology of the crystals. Both changes increase the rate of
Table 2 Typical scale samples in Iranian and international oilfield Moghadaet et al., 2002 Area Persian Gulf (Siri-C) Persian Gulf (Nosrat) Onshore field (Ahwaz) Onshore field (Aghajari) Onshore field (Gachsaran) North Sea West Texas Formation Mishrif Mishrif Bangestan Bangestan Asmari Forties Queen Sand Major scale CaCO3 CaSO4S2H2O CaSO4S2H2O CaCO3 CaCO3 BaSO4, SrSO4 80% SiO2, 20% CaSO4, NaCl 85% CaSO4, CaCO3 84% CaSO4, 10% SiO2, 6% Fe2O3 Minor scale SrSO4 CaCO3 CaCO3 CaSO4S2H2O CaSO4S2H2O CaCO3 CaSO4, Na2SO4 Mg, Fe, Si Si, Al

New Mexico West Texas

Abo San Andres

204

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

permeability decline. The basic structures of crystals formed under dynamic flow conditions and in static solutions are similar, but some faces of the crystals disappear and others become more plate-like when subjected to shear forces. Injection tests on cemented glass bead packs showed a 95.5% reduction in permeability because of the bridging by 2 10-Am calcium carbonate crystals that blocked 33-Am pore throats (Cusack et al., 1987). Thronton and Lorenz (1987) and Crow (1985) showed how sodium and iron hydroxide colloid particles block the pores of quartz sand. They suggested two mechanisms by which the plugging of quartz sand occurs: flocculation/coagulation of the sodium and iron hydroxide phase producing a filter cake near the inlet of the sand packs and colloid-quartz surface interaction by which the particles attach to the quartz surface. In theoretical studies, geochemical flow models were developed (Blout and Dickson, 1973; Khatib, 1994; Khillar and Fogler, 1987; Khillar and Fogler, 1990; Shen and Corsby, 1983; Vitthal and Sharma, 1992; Walsh et al., 1983; Yeboah et al., 1993) for both mineral dissolution and precipitation processes during convective transport. These models were based on the assumption of local thermodynamic equilibrium and negligible dispersion. They did not include solids migration, an assumption, which was severely limiting because many experimental studies demonstrated that precipitated solids could migrate within the porous matrix.

A theory of solids fractional flow was proposed in an attempt to characterize the local permeability reduction by incorporating the dominant pore-plugging mechanism (Wu, 1994). A fractional flow function for each migrating mineral, coupled with the mass conservation of each element and the assumption of local equilibrium between solid and aqueous species, provided a complete description of the reactions during flow through permeable media.

2. Experimental set-up and procedure Experiments were carried out using a test rig, which is schematically shown in Fig. 2. The stainless steel tanks are heated to approximately 90 jC using four band heaters mounted on the outside. A small cooling water coil is installed inside the tanks to allow accurate control of the liquid temperature in the tanks. The thermocouple for determining the temperature inside the tanks is connected to a controller for the band heaters power output. Furthermore, a stirrer is also placed inside the tank to agitate the liquid and to provide a uniform temperature. The peristaltic pump speed can be varied between 0.5 and 55 rpm so that the liquid flow can be accurately adjusted. The maximum design pressure of the pump is 3 bar, which is achieved by using tubing with an internal diameter of 1.6 mm and a wall

Fig. 2. Schematic diagram of the test rig.

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

205

thickness of 1.6 mm. The tubing is made from Marprene II, a material that is resistant to water and mineral oil. The test section, which contains the porous medium, is made of stainless steel pipe with an internal diameter of 32 mm, a wall thickness of 5 mm and a total length of 580 mm. Spiral and longitudinal grooves to accommodate heating wire and thermocouples for measuring the wall temperature have been milled in the outside of the pipe. Bores to insert thermocouples for measuring the sand bed temperature, and opposite to them bores for pressure tapping, have been drilled at locations shown in Fig. 3. The cold-ended thermocoax heating wire is fitted inside hemispherical grooves around the tube at a pitch of 8 mm, to give a heated area of 0.03488 m2. The power (maximum 2000 W) is controlled to maintain the wall temperature below the maximum operating temperature of around 200 jC. The test section has six pressure taps along its length, each of them connected to a separate pressure transducer. Additional perspex test sections have been manufactured and installed for visual observation of the deposition mechanisms.

All transducers have been connected to a computer-controlled data acquisition systems for on-line monitoring and processing of the experimental results. The present investigation is focused on kinetics and mechanisms of CaSO4 and CaCO3 precipitation, rather than on providing quantitative data for oil reservoir application. Therefore, deposition in a well-defined packed bed from spherical glass beads or sand particles with a fixed porosity was chosen. When the bed is fully packed, the porous medium fills the space between the two screens at the ends of the test section. The pore volume of the dry porous medium was then filled with liquid supplied from a burette to determine the porosity as the ratio of the required volume of liquid 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 the medium. The properties of the packing materials and of the investigated fluids are given in Table 3. Preliminary tests were performed to obtain the time after which the bed was stabilized. Distilled water was pumped through the bed for about 1 h to obtain a

Fig. 3. Design of the test section.

206

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217 Table 4 Range of operating parameters in scale formation experiments Inlet temperature Flow rates System pressure Solution viscosity Solution concentration 50 80 jC 25 100 cm3/min 122 kPa 0.7 10 3 1.3 10 3 kg/m s Ca(NO3)2S4H2O 1.8 13.4 g/l Na2SO4 2.6 7.8 g/l Na2CO3 0.8 1.6 g/l 2.5 7.5 g/l CaSO4 CaCO3 0.750 1.5 g/l Solution type A [Ca2 +] = 4.33 g/l [SO] = 2.6 g/l 4 Solution type B [Ca2 +] = 8.67 g/l [SO] = 5.22 g/l 4 Solution type C [Ca2 +] = 13 g/l [SO] = 7.8 g/l 4 Solution type D [Ca2 +] = 1.77 g/l [CO] = 0.8 g/l 3 Solution type E [Ca2 +] = 2.36 g/l [CO] = 1.06 g/l 3 Solution type F [Ca2 +] = 3.54 g/l [CO] = 1.6 g/l 3

homogenous condition. While all operational variables of the system were kept constant, pressure readings were taken at short time intervals until steady state was reached. This was usually achieved after about 20 min. In all experiments, great care was taken to eliminate any air pockets in the bed before any measurements have been taken. The scale formation experiments were performed with aqueous solutions of Na2SO4, Ca(NO3)2S4H2O and Na2CO3. The criteria for selecting these salts were based on the solubility of the salts and the valence of the respective ions. Each test liquid involved two solutions of these salts, one rich in calcium ions and the other rich in sulfate or carbonate ions, which were kept separate until entering the porous medium. Mixing of calcium nitrate (Ca(NO3)2S4H2O) and sodium sulfate (Na2SO4) or sodium carbonate results in the formation of calcium sulfate or calcium carbonate in the porous medium. The range of salt concentrations used in this investigation is given in Table 4. Supersaturated salt solutions have been adjusted for accelerated tests, to allow sufficient variation of operating parameters. Calcium nitrate and sodium sulfate or sodium carbonate solutions of predetermined concentrations were 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. Then the data acquisition system was switched on to record fluid flow rate, temperatures, and pressures. At this time, the supply pumps were switched to the scale forming solutions. The range of investigated operating variables is given in Table 4. At the end of each experiment, the supply pumps were switched to EDTA (ethylene diamine tetra acetic acid) solution and flooding continued for about 20 min. This was to
Table 3 Physical properties of packing material Packing material Sand Particle size range, Am 180 250 250 425 400 500 1000 180 300 250 425 400 600 1000 Mean particle diameter, Am 192 265 410 1000 245 338 480 1000 Porosity 0.3827 0.3830 0.3839 0.3853 0.3787 0.3792 0.3804 0.3825 Permeability, m2 2.14e 11 6.03e 11 1.21e 10 7.27e 10 4.41e 11 7.89e 11 1.59e 10 7.05e 10

Solutions used in the present experimental study

remove from the flow loop all salt forming ions, which could confuse later experiments.

3. Results and discussion 3.1. Clean bed experiments Before carrying out any tests with solutions, it is first necessary to have adequate information regarding the flow mechanisms in a clean medium. When a fluid flows through a porous medium, the pressure drop which develops along the bed in the direction of flow is a function of system geometry, bed voidage and the physical properties of bed and fluid. The operating conditions can result in four distinct flow regimes (Dybbs and Edwards, 1997): Darcy or creeping flow, inertial flow, unsteady laminar flow and chaotic (or turbulent) flow. In the Darcian region, the pressure gradient is proportional to the flow rate and is mathematically expressed by: u K l   Dp Dx 1

Glass beads

The coefficient K for single-phase flow depends only on the geometry of the porous medium. It is called the

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

207

Fig. 4. Variation of pressure drop as a function of axial distance.

specific or absolute permeability of the medium; in the case of single-phase flow, this is abbreviated as permeability. The measurements of pressure drop as a function of axial distance are presented in Fig. 4 for different media and liquid flow rates. As predicted by Eq. (1), there is a linear relationship between the pressure drop and the axial distance in the direction of flow. To investigate the effect of the particle size, experimental data of pressure gradient as a function of flow rate were collected for various particle diameters.

The data are presented in Fig. 5. As can be observed, the pressure gradient increases as the particle diameter decreases. This plot also demonstrates the expected linear relationship between pressure gradient and flow rate. Knowing the viscosity of the liquid saturating the medium, the permeability of each medium can be calculated using the slope of the best-fit straight line through the data presented in Fig. 4 as the ratio of ul/ K in Eq. (1). The calculated permeabilities are also

Fig. 5. Variation of pressure gradient as a function of flow rate.

208

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

given in Table 3. A great deal of effort has been spent to predict the value of K empirically. For the case of simple geometry and unconsolidated porous media, it can be estimated in terms of the geometrical parameters and bed porosity with reasonable accuracy. Various methods for estimation of permeability are documented by Dullien (1992). Values of K for different materials vary widely. Dullien (1992) reported some typical values of permeability of some common porous media. For example, in the case of beds made of particles or fibers, one can introduce an effective average particle or fiber diameter dp. The hydraulic radius theory of Kozeny Carman (Kozney, 1927; Carman, 1937) leads to the relationship:
2 dp2

For packed beds of spherical particles with narrow range of size distribution, Rumpf and Gupte (1971) show that: K /5:5 2 d 5:6 p 4

180a

1 /2 /3

The average effective particle diameter is given by: Z


0 l 3 dp hdp ddp

dp2

0Z
0

l 2 dp hdp ddp

with h (dp) being the density function for the distribution of the diameter dp.

gives a better agreement with the experimental results providing that the considered range of porosity is between 0.35 and 0.67. A comparison of the predictions made by Eqs (2) and (4) is illustrated in Fig. 6. It can be seen that the predictions due to both correlations are in close agreement at lower particle diameters; however, the correlations deviate appreciably in their predictions for larger particles. The measured effect of particle diameter for four particle sizes is presented in Fig. 7. The permeability of the porous medium for singlephase is depends only on the geometry of the medium. As can be observed, the permeability decreases as the particle diameter decreases. As in any fluid flow problem, the range of validity of Darcys law is expressed in terms of a Reynolds number, Re. The Reynolds number is normally defined in terms of a characteristic length of the system in which fluid flow occurs. However, for flow through unconsolidated porous media, it is customary to express Re in terms of a characteristic grain size, and for flow in consolidated porous media in terms

Fig. 6. Comparison of different models for prediction of permeability.

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

209

Fig. 7. Variation of permeability as a function of particle size.

of a mean pore size. In any event, it is generally accepted that Darcys law is applicable to the region where viscous forces are predominant. To investigate the effect of the particle size, experimental data of permeability as a function of modified Reynolds number were collected for various particle diameters. Fig. 8 shows that the permeability of the porous medium in single phase flow is independent of the nature of the fluid and depends on the geometry of the medium.

3.2. Scale formation experiments In this study, the basic knowledge for understanding the kinetics of calcium sulfate and calcium carbonate precipitation has been reviewed. The mechanisms and kinetics of CaSO4 and CaCO3 precipitation have been studied in several investigations (Nancollas et al., 1978; Nancollas and Sawada, 1982; Oddo and Tomson, 1994; Vetter, 1972; Vetter and Farone, 1987; Vetter et al., 1987); however, there

Fig. 8. Variation of permeability as a function of modified Reynolds number.

210

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

is still no clear and reliable kinetic model available to estimate scaling rates under oilfield reservoir conditions. To study these mechanisms, experiments have been designed for demonstrating that an incompatible mixture of solutions can plug an unconsolidated porous medium. This porous medium represents a very small section of the reservoir. The aim was to reproduce local reservoir conditions in this porous medium so that suitable scaling data could be obtained. Supersaturation was hence created by mixing two incompatible brines inside the porous medium, rather than in a supply tank where suspended particles would have been generated due to bulk precipitation. To achieve rapid mixing in the porous medium, a special inlet chamber to the test section was designed to generate complete transverse mixing within the length of the test section. Even though the supersaturation was created by mixing in the test rig, it could represent supersaturation created in the reservoir by any of the other supersaturation generating processes (temperature and concentration). 3.2.1. Calcium sulfate experiments The test rig was designed to operate at temperatures of 50 80 jC, i.e. reasonable temperatures for Iranian oil fields. The flow rates used in these experiments range from 25 to 100 cm3/min, corresponding to solution velocities in the range between 0.05 and

0.2 cm/s. These velocities were intended to reproduce brines flowing through the rock immediately adjacent to producing wells, where scale is expected to be most damaging. The solution velocities used are rather high for Iranian oil fields. During each run, the pressure drop across the test section was recorded continuously. The pressure drop increased during the experiments only when a supersaturated solution was flowing through the test section. This confirms that the increase is caused by scale formation. Typical results for pressure drop and permeability reduction obtained from the tests are presented in Figs. 9 11. The change of permeability and the pattern it follows are the most significant pieces of information to be gained from the experimental study. All the transient permeability plots obtained from the experiments show a more or less concave trend. In the following, the results for various temperatures, concentrations, and flow rates are discussed individually: Temperature has a significant influence on solubility and crystal growth of calcium sulfate. To study its effect on the permeability reduction, a number of tests were carried out, where flow rate and concentration of the injected solution were kept constant and the temperature was varied. These tests were carried out at constant injection rates of 25 cm3/min, 4.33 g/l calcium ion and 2.6 g/l sulfate ion concentration, respectively,

Fig. 9. Variation of pressure drop as a function of time.

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

211

Fig. 10. Variation of permeability ratio as a function of time showing effect of temperature.

and temperatures of 50 and 80 jC. Fig. 10 shows the variation of permeability reduction with time at different temperatures. The permeability decline is more rapid at higher temperature, since the rate of precipitation and the supersaturation both increase with temperature. To investigate the effect of flow rate on permeability reduction, a set of tests were performed in which the temperature and concentration of solutions were

kept constant while the flow rate was varied. These tests were carried out at a temperature of 70 jC, 4.33 g/l calcium ion, and 2.6 g/l sulfate ion concentration respectively, and flow rates of 50 and 100 cm3/min. Fig. 11 shows variation of the permeability decline as a function of time. From these figures, the permeability decline of the porous medium is evident, even at such low flow rates. The overall permeability of the porous

Fig. 11. Variation of permeability ratio as a function of time showing effect of flow rate.

212

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

Fig. 12. Variation of permeability ratio as a function of flow rate.

medium decreased to between 35% and 61% of the initial permeability, depending on the flow rate used. At higher flow rates more calcium and sulfate ions will enter the porous medium in a given interval of time, hence providing more material for deposition. However, the decline in permeability ratio was not linear with flow rate, due to the competing effects of deposition and shear-related removal.

Fig. 12 shows variation of permeability decline as a function of flow rate. In a more general way, the plot of permeability ratio versus logarithm of the fluid flow rate (Fig. 13) was found to be linear for different injection rates. The following relationship was found:       K K Q2 mln 5 K i Q2 Ki Q1 Q1

Fig. 13. Exponential relationship between flow rate and permeability ratio.

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

213

Fig. 14. Variation of permeability ratio as a function of time showing the effect of concentration.

The values of m can be extracted from Fig. 13; they can be correlated by an exponential function of pore volume injected as follows: m 1 eaPVI Substituting the values of m into Eq. (5)       K K Q2 aPVI 1e ln Ki Q2 Ki Q1 Q1 6 7

The above simple relationships may be used to predict the permeability decline in a similar material at any

new flow rate. From Figs. 12 and 13, it is observed that the permeability decline due to scale formation is not only a function of pore volume injected (time), but also depends on the flow velocity. At higher flow velocities, the process of scaling increases and therefore, the overall permeability becomes lower. A number of tests were carried out to study the effect of solution concentration on the permeability reduction. These tests were carried out at an injection rate of 50 cm3/min, with three different solution

Fig. 15. Variation of permeability ratio as a function of time showing the effect of temperature.

214

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

Fig. 16. Variation of permeability ratio as a function of time showing the effect of flow rate.

concentrations (solution types A, B and C, see Table 4). Fig. 14 shows the variation in permeability decline with time for different concentrations. When the concentration of the solution (i.e. supersaturation) is increasing, plugging and hence permeability loss occurs more rapidly. Calcium carbonate experiments: The procedure for these experiments was identical to that used for calcium sulfate. Since for both salts, similar mechanisms are responsible for the increase in pressure drop

and the decrease in permeability, the same trends are observed and the explanations given above apply to the calcium carbonate results, as well. Figs. 15 and 16 show that the rate of permeability reduction due to CaCO3 deposition becomes faster with increasing temperature and flow rate. As for calcium sulfate, the plot of permeability ratio versus log of flow rate was found to be linear for different CaCO3 injection rates. The value of parameter m for Eq. (5) can be extracted from Fig. 17.

Fig. 17. Exponential relationship between flow rate and permeability ratio.

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

215

Fig. 18. Exponential relationship between flow rate and permeability ratio.

Fig. 18 shows the variation of permeability ratio as function of the logarithm of the flow rate for both, calcium sulfate and calcium carbonate scale formation. The permeability decline due to formation of calcium carbonate scale is greater than for calcium sulfate, for the given experimental conditions.

has been suggested and verified against the present data. Nomenclature A cross-sectional area, m2 C total amount of ionic species in solution, M Cs mass solid concentration, kg/m3 dc test section diameter, m dp particle diameter, Am dp pressure drop, N/m2 dp/dx pressure gradient in x direction, N/m3 EDTA ethylene diamine tetra acetic acid I ionic strength, mol K permeability, m2 KI initial permeability, m2 m mass fraction, slope or rate of scale deposited, kg/m3 overall section measuring during total length of porous medium, section 1 until section 6 (41.5 cm) p pressure, N/m2 Dp/l pressure gradient, Pa/m Dp pressure drop, Pa Q0 cumulative volume of fluid injected expressed in terms of the initial undamaged pore volume, m3 Q volumetric flow rate, m3/s Rem modified Reynolds number defined as dpqu//l(1 /)

4. Conclusions 1. Permeability decline caused by scale formation in the porous bed ranged from less than 30% to more than 90% of the initial permeability, depending on solution composition, initial permeability, temperature, flow rate and solution injection period. 2. The pattern of permeability decline in a porous medium due to scaling injection was characterized by a steep initial decline which gradually slowed down to a lower, but often still significant, constant damage rate. The initial steepness of these curves generally decreased with increasing distance from the point of mixing of the incompatible solutions. 3. Several factors influencing scale formation have been examined. Both increasing temperature and flow rate have a detrimental effect on the permeability reduction. 4. A simple correlation between flow rate, initial permeability and permeability reduction with time

216

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217

Ref Reo T Ts t time of u x

Reynolds number defined as dpquf/l Reynolds number defined as Dqu/l temperature, jC, K solids concentration in flow stream, kg/m3 time, min, s injection period of fluid injection in test section, Qo/Q velocity of flowing phase through porous medium, m/s distance from inlet face of test section or core, m

Greek Symbols / porosity /I initial porosity lf fluid viscosity, kg/m s D gradient Subscripts hr hour i initial or inlet l liquid L length m matrix, mean and medium PV pore volume PVI pore volume injection, permeability variation index sph sphere TC thermocouple Superscripts j degree f notation V feet W inch

References
Atkinson, G., Raju, K., Howell, R.W., 1991. The thermodynamics of scale prediction paper NO. 21021. Paper presented at Society of Petroleum Engineers International Symposium on Oilfield Chemistry, Anaheim, CA, 20 22 Feb., No. 21021. Bezmer, C., Bauer, K.A., 1969. Prevention of carbonate scale depositiona well packing technique with controlled solubility phosphates. Journal of Petroleum Technology 246, 505 514. Blout, C.W., Dickson, F.W., 1973. Gypsum anhydrite equilibrium in systems CaSO4 and CaSO4 NaCl H2O. American Mineralogist 58, 1746.

Carman, P.C., 1937. Fluid flow through granular beds. Transactions of the Institution of Chemical Engineers (London) 15, 150. Cowan, J.C., Weintritt, D.J., 1976. Water Formed Scale Deposits. Gulf Publishing, Houston, p. 586. Crow, C.W., 1985. Evaluation of agents for preventing precipitation of ferric hydroxide from spent treating acid. Journal of Petroleum Technology, 691 695 (Apr.). Cusack, F., Mc-Kinley, V.L., Lappin-Scott, H.M., Brown, D.R., Clementz, D.M., Costerton, J.W., 1987. Diagnosis and removal of microbial/fines plugging in water injection wells. Paper presented at 62nd Annual Technical Conference and Exhibition of the Society of Petroleum Engineers held in Dallas, TX, 27 30 Sept., No. 16907. Dullien, F.A.L., 1992. Porous Media: Fluid Transport and Pore Structure. Academic Press, New York. Dybbs, A., Edwards, R.V., 1997. Department of fluid, thermal and aerospace science report. FTAS/Tr (Case Western Reserve University) No. 75-117, Workshop on Heat and Mass Transfer in Porous Media (PB-252-387). NTIS, Springfield, VA, p. 228. El-Hattab, M.I., 1982. Gupcos experiences with treating gulf of squeeze seawater for water flooding the El-Morgan oil field. Journal of Petroeum Technology, 1449 1460 (July). Khatib, Z.I., 1994. Prediction of formation damage due to suspended solids: modelling approach of filter cake build-up in injectors. Society of Petroleum Engineers Paper No. 28488, Society of Petroleum Engineer 69th Annual Technical Conference and Exhibition, New Orleans, LA, Sep. 25 28. Khillar, K.C., Fogler, H.C., 1987. Colloidally induced fines migration in porous media. Reviews in Chemical Engineering 4 (1 and 2), 41 108. Khillar, K.C., Fogler, H.C., 1990. Colloidally and hydrodynamically induced fines migration in porous media. In: Cheremisinoff, N.P. (Ed.), Surface and Ground Water Flow Phenomena. Encyclopaedia of Fluid Mechanics, vol. 10, pp. 623 663. Chap. 19. Kozney, J., 1927. Uber kapillare Leitung des Wassers im Boden. Bericht Wiener Akademia 136, 271. Krueger, R.F., 1986. An overview of formation damage and well productivity in oil field operation. Journal of Petroleum Technology, 131 152 (Feb.). Krumine, P.H., Mayer, E.H., Brock, G.E., 1985. Scale formation during alkaline flooding. Journal of Petroleum Technology, 1466 1474 (Aug.). Lasetr, R.M., Grander, T.R., Glasscock, F.M., 1968. Scale deposits are controlled now with liquid inhibitors. Oil and Gas Journal, 63 88 (Jan. 15). Leon, J.A., Scott, E.M., 1987. Characterization and control of formation damage during water flooding at high clay content reservoir. Paper presented at Society of Petroleum Engineers Production Operation Symposium, Oklahoma City, 8 10 Mar, No. 16234. Lindlof, J.C., Stoffer, K.G., 1983. A case study of seawater injection incompatibility. Journal of Petroleum Technology, 1256 1262 (July). Liu, S.T., Nancollas, G.H., 1971. The kinetics of dissolution of calcium sulfate dihydrate. Journal of Inorganic and Nuclear Chemistry 33, 2311. Mitchell, R.W., Grist, D.M., Boyle, M.T., 1980. Chemical treat-

J. Moghadasi et al. / Journal of Petroleum Science and Engineering 43 (2004) 201217 ments associated with North Sea projects. Journal of Petroleum Technology, 904 912. Moghadasi, J., Jamialahmadi, M., Muller-Steinhagen, H., Sharif, A., Izadpanah, M.R., 2002. Formation damage in Iranian oilfield. Paper presented at Society of Petroleum Engineers Formation Damage Control International Symposium and Exhibition, Lafayette, LA, USA, 20 21 Feb, No. 80460. Moghadasi, J., Jamialahmadi, M., Muller-Steinhagen, H., Sharif, A., Ghalambor, A., Izadpanah, M.R., Motaie, E., 2003. Scale formation in Iranian oil reservoir and production equipment during water injection. Paper Society of Petroleum Engineers 80406 Presented at the 5th International Oilfield Scale Symposium and Exhibition held in Aberdeen, UK, 29 30 January. Nancollas, G.H., Campbell, J.R., 1969. The crystallisation and dissolution of strontium sulfate in aqueous solution. Journal of Physical Chemistry 23 (6), 1735 1740. Nancollas, G.H., Liu, S.T., 1975. Crystal growth and dissolution of barium sulfate. Society of Petroleum Engineers Journal, 509 516. Nancollas, G.H., Sawada, K., 1982. Formation of scale calcium carbonate polymorphs: the influence of magnesium ion and inhibitor. Journal of Petroleum Technology, 645 652 (March). Nancollas, G.H., Eralp, A.E., Gill, J.S., 1978. Kinetics of growth of calcium sulfate scale formation: a kinetic approach. Society of Petroleum Engineers Journal, 133 138. Oddo, J.E., Tomson, M.B., 1994. Why scale forms and how to predict it. Paper presented at Society of Petroleum Engineers Symp. on Production Operations, Oklahoma City, Ok, 7 9 April, No. 21710. Pang, R., Sharma, M.M., 1994. A model for predicting injectivity decline in water injection wells. Society of Petroleum Engineers Paper No. 28489, Society of Petroleum Engineers 69th Annual Conference and Exhibition, New Orleans, La., Sept. 25 28. Rumpf, H., Gupte, A.R., 1971. Die Porositatsfunktion in empiri schen Gleichungen fur den Durchstro mungswiderstand im Bereich Re V 1. Chemieingenieur Technik 43, 376. Shen, J., Corsby, C.C., 1983. Insight into strontium and calcium sulfate scaling mechanisms in oil producing wells. Journal of Petroleum Technology, 1249 1255 (July). Singley, J.E., 1985. Corrosion Prevention and Control in Water Treatment and Supply Systems. Noyes Publication, Houston. Thronton, S.D., Lorenz, P.B., 1987. Role of silicate and aluminate

217

ions in the reaction of sodium hydroxide with reservoir minerals. Paper Society of Petroleum Engineers 16277 Presented at the International Symposium on Oilfield Chemistry, San Antonio, TX, Feb. 4 6. Todd, A.C., Yuan, M.D., 1991. Prediction of sulfate scaling tendency in oilfield operations. Society of Petroleum Engineers Production Engineering Journal, 63 72 (Feb.). Valone, W.F., Skillern, K.R., 1982. An improved technique for predicting the severity of calcium carbonate. Paper presented at Society of Petroleum Engineers 6th International Symp. on Oil Field and Geothermal Chemistry, Dallas, TX, 25 27 Jan., No.10549, pp. 11 32. Paper No. 10549. Vetter, O.J., 1972. Oilfield scalecan we handle it? Journal of Petroleum Technology, 1402 1408 (Dec.). Vetter, O.J., Farone, W.A., 1987. Calcium carbonate scale in oilfield operations. Society of Petroleum Engineers paper No. 16908 presented at the 62nd Annual Technical Conference and Exhibition held in Dallas, September 27 30. Vetter, O.J., Kandarpa, V., Harvaka, A., 1982. Predication of the scale problems due to injection of incompatible waters. Journal of Petroleum Technology, 273 284 (Feb.). Vetter, O.J., Farone, W.A., Veith, E., Lankford, S., 1987. Calcium carbonate scale considerations: a practical approach. Society of Petroleum Engineers paper No. 17009 presented in the International Symposium of Production Technology held in Lubbock, November 16 17. Vitthal, S., Sharma, M.M., 1992. Dynamics model for particle deposition and build up in granular media. Journal of Colloidal and Interface Science 153 (2), 314 336 (Oct.). Walsh, M.P., Rouse, B.A., Senol, N.N., Pope, G.A., Lake, L.W., 1983. Chemical interactions of aluminium citrate solutions with formation minerals. Paper presented at Society of Petroleum Engineers International Symposium on Oilfield and Geothermal Chemistry, Denver, CO, 1 3 June, No. 11799. Wu, G.G., 1994. A PC-windows-based program for geostatistical modelling application. Paper presented at the Society of Petroleum Engineers Petroleum Computer Conference, Dallas, TX, 31 July 3 August, No. 28242. Yeboah, Y.D., Somuah, S.K., Saeed, M.R., 1993. A new and reliable model for predicting oilfield scale formation. Paper presented at Society of Petroleum Engineers International Symposium on Oilfield Chemistry, New Orleans, LA, No. 25166.

You might also like