Sie sind auf Seite 1von 9

SPE 79695 Coupled Simulation of Reservoir Flow, Geomechanics, and Formation Plugging With Application to High-Rate Produced Water

Reinjection

R.C. Bachman, TAURUS Reservoir Solutions Ltd., T.G. Harding, Nexen Inc., A. (Tony) Settari, U. of Calgary and D.A. Walters, TAURUS Reservoir Solutions Ltd.
Copyright 2003, Society of Petroleum Engineers Inc. This paper was prepared for presentation at the SPE Reservoir Simulation Symposium held in Houston, Texas, U.S.A., 35 February 2003. This paper was selected for presentation by an SPE Program Committee following review of information contained in an abstract submitted by the author(s). Contents of the paper, as presented, have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material, as presented, does not necessarily reflect any position of the Society of Petroleum Engineers, its officers, or members. Papers presented at SPE meetings are subject to publication review by Editorial Committees of the Society of Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O. Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

including multiphase flow, geomechanics (stress changes), formation plugging, and fracture mechanics. This paper describes the formulation and numerical implementation of a model, which treats all of the above phenomena. The model is an extension of a previous coupled reservoir and geomechanics model3,4, combined with a dynamic fracture propagation feature and permeability reduction model. There are several possible approaches to fracture propagation modeling, which will be discussed in detail. The plugging mechanics is based on a simple, yet realistic model that can be easily implemented in any conventional simulator. The formulation described here has been implemented in the GEOSIM modeling system and also in the Open Eclipse environment. The software has been used to model the PWRI injection in the Masila block in Yemen, operated by Nexen Inc. Although the details of the engineering study are beyond the scope of this paper, the methodology of conducting such studies and the consequences of the plugging mechanics for history matching will be presented in detail. The model shows in particular the importance of the coupling between the plugging, which generates increased pressure gradients, the associated poroelastic and thermoelastic stress changes, and the resulting fracture propagation pressure. Formulation of the model The coupled model consists of four main components: Fluid flow model Deformation and stress model Dynamic fracture model Permeability damage model Because the formulation of the first two components has been described previously3,4, this Section will give only a brief overview and highlight the couplings between them, followed by a detailed discussion of the fracturing and the damage model in the following Sections. Reservoir flow. This part of the system is a conventional 3-dimensional 3phase black oil simulator. This model is the host or master

Abstract Produced water re-injection at high rates presents a coupled problem of reservoir flow, formation damage, stress alteration around the injector, and fracture propagation. Accurate prediction of permeability changes, fracture propagation pressure, and fracture dimensions is required for minimization of disposal costs and design of surface equipment. The paper presents the formulation and numerical implementation of a coupled reservoir, damage and geomechanical model which includes the above couplings. Details of the model are first described. A simple empirical damage model, calibrated to field data is then presented. Finally, application of the complete model to high rate reinjection in the Masila Block in Yemen is presented. The model predictions show that it is feasible to sustain over 100,000 BWPD in a single Masila disposal well by injecting above fracture pressure. Introduction Oil production operations often produce large volumes of water and high rate produced water re-injection (PWRI) is usually the best method of water disposal. However, injection wells can experience large reduction of injectivity due to plugging caused by solids and oil in water1,2. In particular, the combination of total suspended solids (TSS) and oil-in-water (OIW) is particularly damaging2 and can cause equivalent skins on the order of 200 or more. Consequently, injection pressures increase with time and induced fracturing may take place. The prediction of injectivity and fracture propagation in injectors experiencing damage is a complex coupled problem

R.C. BACHMAN, T. HARDING, A. SETTARI AND D.A. WALTERS

SPE79695

for the other components. In the actual implementation, of the software, the host reservoir model can be either Eclipse 100 or the DE&S thermal reservoir simulator TERASIM. Deformation/stress model The model used is a continuum finite element code which solves the classical poro and thermoelasticity equations for nonlinear elasticity in an incremental fashion. The model also has elasto-plastic capabilities (discussed in Ref. 3) which have not been used in this work. In typical geomechanical applications, the main couplings between the flow and stress are through the changes in porosity (pore volume coupling) and through stress dependent permeability (flow properties coupling). The first is dominant in compaction problems5, while the second is important in problems such as waterfloods in jointed media6. In PWRI problems, both of these couplings are of minor importance until formation failure is induced around the fracture. However, there are additional couplings arising from fracturing and damage mechanisms: Permeability reduction due to damage can be very large (especially close to the injector), and will completely overshadow any stress-dependent changes. Stress changes around the injector due to pressure and temperature changes cause time-dependent changes in fracture initiation and propagation pressure. Because the pressure gradients are a strong function of the damage, there is a strong coupling between the damage, time of the start of fracturing, and fracturing pressures.

fracture mechanics is significant. Therefore, the ultimate PWRI fracturing model consists of three fully coupled components: 1) Fracture mechanics model (computing fracture geometry, flow and heat transfer inside the crack), 2) reservoir model (computing flow, heat transfer and damage in the reservoir) and 3) geomechanics (computing stress-strain response of the reservoir and its surroundings to pressure and temperature changes, and loading on fracture face). Some of the important couplings between the components are: Pressure in the fracture is a boundary pressure for reservoir flow (leak-off) Fracture width and pressure are equal, respectively, to the displacement and normal stress on the crack boundary in the stress model Pressure and temperature in the reservoir are loads for the stress solution Effective stress and volumetric strain determine reservoir permeability and porosity

The modeling approach can be either to formulate the entire problem in a fully coupled manner, or in a modular fashion. In a fully coupled model, the couplings become internal compatibility conditions and are satisfied automatically, but require internal iterative process. In the modular approach used in this work (see Fig. 2 of Ref. 4), the compatibility conditions can be satisfied by external iteration between the modules, or approximated. We will now describe two existing methods of implementing fracture modeling in the context of a modular system and discuss the outstanding issues for rigorous coupling. Both methods are based on the assumption that the primary interest in the modeling is to predict well productivity or injectivity under fracturing conditions, and the fracture geometry details are not as critical. Partially coupled approach The first is an extension of the partially coupled concept for modeling stimulation treatments (Settari et al. 1990). The fracture propagation is pre-computed using an uncoupled fracture mechanics model (such as the models discussed under a) above). For the coupling with reservoir, the fracture must lie in a grid plane of the flow model and it is represented in the flow model by increased transmissibilities, which are calculated from the local fracture width. The stress model shares the mesh with the flow model; but the fracture width is typically not used as a displacement boundary condition in fracture plane. Since the transmissibilities are re-computed in time, the model can represent the dynamic process of fracture growth, or the propped or acidized fracture after the treatment (essentially static except for the conductivity changes due to stress). This approach also ignores the effect of created fracture storage on pressure (this effect is however significant only when the fluid efficiency is high). Since the leak-off during the fracture propagation calculations is computed independently, there is no guarantee that it will be correct. This will show up as a mismatch in computed injection pressures from the fracture mechanics and reservoir model: if

Modeling of induced fracturing in geomechanics Fracture modeling can be approached from either the fracture mechanics side, or the reservoir flow side. a) The conventional hydraulic fracturing models (intended primarily for stimulation treatments) focus on details of fracture geometry and other fracturing physics, and decouple the reservoir flow by the use of analytical leak-off models. It is well known that this approach becomes inaccurate as the leak-off increases. The analytical reservoir treatment can be improved considerably7 and incorporate the plugging effects8,9 but ultimately such models are limited in generality. b) The alternative approach is to build a model of fracturing into a reservoir simulator to treat the leak-off implicitly. Such models have been known since the 1980s. Early examples10,11 coupled 2-D analytical or pseudo-3D fracture geometry models with 3-D, 2-phase, thermal reservoir models. However, the coupling with stress was still not included and consequently the fracture propagation pressure had to be specified. For PWRI problems, the stress coupling and its effect on

SPE 95695

COUPLED SIMULATION OF RESERVOIR FLOW, GEOMECHANICS AND FORMATION PLUGGING

the predicted fracture growth rate is too small, coupling this fracture into the reservoir model will produce pressures which are too high and vice versa. Similarly, the poro and thermoelastic back stresses (also referred to as back stresses) are computed independently in the fracture model, and do not necessarily agree with the stresses on the fracture face computed by the stress model. In spite of these limitations, this method has been used extensively and can be applied successfully to a number of problems if the models are tuned to produce the same pressure history13,14. Fully coupled approach with simplified fracture mechanics The second approach is similar to that presented recently for modeling waterfracs15. The idea is to use stress-dependent permeability functions, which can represent the flow behavior of the fracture in a fully coupled manner. In the potential fracture plane (normal to the minimum effective stress), the permeability (or directly the flow transmissibility) is increased by several orders of magnitude as the effective stress normal to the fracture decreases. A typical function (permeability multiplier) is shown in Fig. 1.
100000

from the opening will be small. Second, the fracture volume is not represented. Again, this will not be a problem in high leakoff situation, but this aspect can be modeled rigorously by effective stress dependent porosity of the fracture blocks. Finally, the tip growth variables need to be introduced. This may be important in hard rock, if fracture propagates through a layered stress system (as it is usually the case), vertical growth through stress barriers may be poorly represented if the fracture opening is not included in the stress model. In the example of Fig. 2, where high stress layer is present within the perforated interval, the approach may produce two separate fractures while the true solution may be a single fracture. Therefore the model will tend to exaggerate the confinement of the fractures.

Confining stress

10000

Transmissibility multiplier

1000

Increasing net pressure in fracture

Fracture mechanics model

Stress dependent k model

100

Fig. 2 Possible fracture geometry differences On balance, our experience indicates that for PWRI fracturing, the reservoir flow and coupling with stress are more important than the fracture mechanics features, and therefore the coupled method was used in this work. Its most significant advantage for field applications is that it can model fractures in different directions (in Cartesian as well as in Corner Point Geometry grid), and within local grid refinement, as shown in Fig. 3.

10

1 -600

-400

-200

200

400

600

Effective stress normal to fracture

Fig. 1 Typical stress-dependent transmissibility functions to represent fracture in the flow model The shape of the curve can be related to the fracture width versus net pressure (i.e., stiffness) and its position to net pressure in the fracture. Imposing a maximum is necessary to maintain the stability of the model. The coupled reservoir and stress model is then run without any reference to a fracture mechanics model, and the region where the flow transmissibilities have reached large values is deemed to represent the fracture. This approach provides implicit coupling between reservoir stress and fracture propagation pressure (i.e., the back stress is implicit), However, the model lacks the fracture mechanics features. First, the crack opening is not included as a boundary condition on the stress model. In soft formations with small modulus and low fracture net pressure, the additional stress

Fig. 3 Possible fracture representations using the transmissibility modifier method Further work is ongoing to remove some of its limitations and to make it a fully coupled solution for general applications.

R.C. BACHMAN, T. HARDING, A. SETTARI AND D.A. WALTERS

SPE79695

The damage model Rigorous modeling of formation damage involves solving the equations for particle transport and entrainment in porous media. While such models have been developed16,17, they are currently too complex for full-field application and require parameters directly describing the physics, which are usually not readily available. For modeling purposes, it is desirable to have a simple model with few parameters, which can be calibrated directly against field injection data. Several authors postulated a model in which the permeability reduction at a given point in the media is expressed in a form:
18,19,20

Damage model with constant n=1, Rm in=0


1 0.9

(k) damaged/(k)initial

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.1 1 10 100

alpha=0.01 alpha=0.1 alpha=1.0

Volume througput/Area

1 k / k0 = (1 + )

Fig. 4 Effect of on damage function


Damage model with constant =0.5, Rm in=0
1 0.9

(1)

(k) damaged/(k)initial

where is some measure of the concentration of the particles at that location. We now make an assumption that can be related to the amount of the water that passed through this location. In a one-dimensional setting, this is simply

n=1.0 n=0.6 n=1.4

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.1 1 10

(t ) V (t ) / A = Q(t )dt
0

(2)

In a finite difference form, at the end of time step K,

100

(t K ) V K / A = Q n t n
n =1

Volume througput/Area

(3)

Fig. 5 Effect of exponent n on damage function


Damage model with constant alpha=0.5, n=1
1 0.9

(k) damaged/(k)initial

where Qn is the computed water flow rate through the area A during the time step n. Based on testing on field data described later, the above formula was further generalized to the following:

Rmin=0.0 Rmin=0.1 Rmin=0.2

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1

k / k 0 Rmin 1 = 1 Rmin (1 + (V / A) n )

(4)

where , n and Rmin are the three parameters of the model. The parameter represents the intensity of the damage and is primarily related to the water quality (TSS and OIW) in combination with reservoir permeability. Increasing accelerates damage as shown on Fig. 4. The exponent n changes the shape of the damage curve as shown in Fig. 5. The factor Rmin was introduced because of field evidence that the damaged permeability does not decrease to zero but reaches an asymptotic minimum value, which in Eqn. (4) becomes kmin = k0 Rmin (see Fig. 6). Laboratory tests on cores21 and field tests in the Masila Block22 suggest that there is also a dependence of damage on flow velocity. Fines migration testing in the lab has indicated a dependence of permeability on flow velocity. In the field, there is little or no damage observed during production tests conducted immediately after drilling and completing disposal wells nor during low rate injection tests, while damage is definitely observed even in short duration high rate injection tests. This aspect is being investigated further.

0 0.1 1 10 100

Volume througput/Area

Fig. 6 Effect of residual Rmin on damage function Implementation Equation (4) was extended to 3-D flow and implemented in the flow calculation via another set of transmissibility multipliers (which are cumulative to those arising from fracture propagation). It should be noted that, in the use of Eqn. (4), a distinction must be made between in-situ reservoir water flow and injected water flow, because it is assumed that the flow of the in-situ water does not create damage. Tracking the difference can be accomplished in two ways: a) Defining two water components in the model (e.g., by using the oil component for in-situ water and water component for injected water). This is often sufficient as the injection usually takes place into a water zone. b) Using a tracer tracking capability in the reservoir model. This is more general as it retains all model capabilities and

SPE 95695

COUPLED SIMULATION OF RESERVOIR FLOW, GEOMECHANICS AND FORMATION PLUGGING

allows injecting several produced water types with distinct plugging properties. The first method is used in the TERASIM based model, and the second in the Eclipse based one. Validation Gulf of Mexico data As an example, consider the data for the Gulf of Mexico (GOM) wells reported in Ref. 1. All wells were limited to injecting below fracture pressure, and temperature effects are small. This allows the testing of the damage model in a simple setting without stress or fracture coupling. The reservoir data is found in Ref. 1. Well A39 started injecting at 5000 BWPD but the injectivity decreased to below 1000 BWPD in less than a year. The decline was matched with the overall value of = 0.12 (1/ft)n and n=1 as shown in Fig. 7. The value was decreased temporarily at early times to account for the acid jobs. At late times, damage was limited by Rmin = 0.0002.
Bullwinkle A39
9000 8800 8600 8400 8000

quality have been obtained for all other wells. The effect of the acid treatments performed in some of the wells can be modeled by the removal of the damage around the well. The method of treating the damage, although very simplified, is remarkably realistic. Due to its empirical nature, the damage parameters must be obtained by history matching of field data, or laboratory experiments. However, when the field injection involves fracturing, the matching process is more complex, as will be shown next on the example of the Haru 4 well of the Masila project. Application high rate injection in the Masila Block The Masila injection project Canadian Nexen Petroleum Yemen Ltd. operates the Masila project in the Republic of Yemen on behalf of its partners Occidental Petroleum Ltd., Consolidated Contractors Company S.A.L. and the Government of the Republic of Yemen. Oil production comes mainly from the high porosity and permeability Upper Qishn sandstones of Lower Cretaceous age. Currently, the operation produces 230,000 BOPD and over 1,000,000 BWPD. The produced water is reinjected at matrix injection pressures mainly into the Upper Qishn section below the original oil-water contact in each field. The target injection horizon is the S2/S3 zone, which is connected to the large regional aquifer, which underlies the producing areas. Despite the high quality of the reservoirs, injectivity problems have hampered the project since inception and have prompted the investigation of causes of these problems and the development of remedies. Further details of the Masila operation and the analysis of water injectivity in the laboratory and field have been reported earlier21, 22. Produced water is currently reinjected into 24 vertical plus 4 horizontal wells. These wells experience poor initial injectivity that decreases further as a result of impurities in the water that are impractical to remove through well head filtration. Well acidizing or proppant fracturing the injectors provided only temporary increases in injectivity. Additional disposal wells will be needed as water production continues to increase. To meet future needs, Nexen plans to drill 4 to 6 injection wells to allow for the additional wastewater disposal of approximately 500,000 BWPD. The present work was undertaken to evaluate high rate produced water reinjection above formation parting pressure. A simulation study was conducted (and is currently being updated) to provide insight into the damage mechanism in the Qishn sands, determine the injectivity and required well spacing at fracture conditions, and ultimately predict the surface injection pressures at various target rates. In this way, the injection project can be optimized with respect to number of wells needed versus the cost for the surface equipment. In this paper, we only present the parts of this work that illustrate the particular features of the coupled modeling and its results. The results of the engineering study including the details of the geomechanics will be published in a future article.

BHP data Inj rate data Model match

7000 6000 5000 4000 3000 2000 1000 0 450

8200 8000 7800 7600 7400 0 50 100 150 200 250 300 350 400

time (days)

Fig. 7 Match of productivity decline GOM Well A39 Well A42 has similar history and its match, shown in Fig. 8, was obtained with the same parameters as for A39. The well
Bullwinkle A42
9000 8800 8600 8400 8000

BHP data Inj rate data Model match

7000 6000 5000 4000 3000 2000 1000 0 350

8200 8000 7800 7600 7400 0 50 100 150 200 250 300

time (days)

Fig. 8 Match of productivity decline GOM Well A42 had a high initial skin, which was accounted for by initial permeability reduction close to the well. Matches of similar

Inj Rate (BWPD)

BHP (psia)

Inj Rate (BWPD)

BHP (psia)

R.C. BACHMAN, T. HARDING, A. SETTARI AND D.A. WALTERS

SPE79695

Calibrating the model Two wells were used to calibrate the model: Camaal 30 and Haru 4. The Camaal 30 well was completed in the S2 zone, with net pay of 102 ft and permeability of 1307 md. Injection began June 26, 1999 and continued with some interruptions until Feb.4, 2001. During that time a number of workovers took place, including filter changes, acid jobs and a propped frac job. Conventional analysis using Hall plots and equivalent skin calculations clearly showed the temporary nature of injectivity improvement from these interventions. History matching established the damage parameters, but because Camaal 30 is a low injectivity well, more emphasis was placed on the Haru 4 calibration. The Haru 4 well is more representative of future injectors. The well is completed in water bearing S2 zone with net pay of 96 ft and average permeability of 4300 md. The injection history from February 1, 1999 to January 6, 2001 is in Fig. 9, together with an interpreted skin due to damage, assuming radial flow.
Haru-4 - Injection Rate and BHP and interpreted damage skin 3000 90000

poroelastic stress change pe, which increases fracture pressure. c) Friction pressure loss in the fracture pfric (part of the net pressure), directly related to fracture conductivity. In PWRI injection, plugging occurs inside the fracture as well and can significantly increase injection pressures. If the fracture conductivity is finite, pfric increases significantly with fracture length xf. Thus, the observed injection pressure pf during fracturing can be written in a simplified manner as pf = c - T (T) + pe (dam) + pfric (dam, xf) (5)

which illustrates the competing nature of these parameters. For Haru 4, thermal expansion coefficient was not measured at the time. Based on literature data, a median value of aL = 0.648 x 10-5 1/0F and a high value of 1.6 x 10-5 1/0F were selected. The cases of low and high fracture conductivity were constructed by choosing the maximum of the fracture multipliers shown in Fig. 1 as 104 and 105. The damage parameters were then varied to obtain a match for each case. Fig. 10 shows the comparison of the pressure match for three cases: Case 1: Median aL = 0.648 x 10-5, low frac conductivity Case 2: High aL = 1.6 x 10-5, low frac conductivity Case 3: High aL = 1.6 x 10-5, high frac conductivity
2700 100000

2500 BHP (psia), damage skin x 10

80000 Injection Rate (stb/d)

2000

70000

1500

60000

1000 BHP skin from damage Injection Rate 0 0 100 200 300 400 Time (days) 500 600 700

50000

2500 BHP, data and simulated (psia)

90000 Injection rate (BWPD)

500

40000

2300

80000

30000 800

2100

70000

Fig. 9 Field data for Haru 4 and interpreted damage skin It was modeled with a single layer reservoir model coupled with a 5-layer geomechanical model, using a Cartesian grid. Permeability damage model was used and fracture propagation (if predicted by the model) could take place. Highly refined grid was used in the potential fracture path and the model was set up to act in an infinite manner. Water was injected at a temperature of T = 27 deg F below reservoir temperature in all cases. A total of 701 days of injection was modeled. It became quickly apparent that the couplings present in the physical system allow non-unique interpretation, if the pressure is the only data matched. Three major factors were identified: a) Thermal expansion coefficient aL. Its value controls the thermal stress component T which is added to the far-field confining stress c . This in turn changes the fracture initiation and propagation pressure. b) Damage strength (parameters , n and Rmin). Increased damage accelerates fracturing, but also increases the

1900

Observed BHP case 1 Case 2 Case 3 inj rate

60000

1700

50000

1500 0 100 200 300 400 time (days) 500 600 700

40000 800

Fig. 10 Comparison of pressure matches with three sets of coupling parameters These matches required different damage parameters as shown in Table 1. Case 1 had the lowest amount of thermal stress and therefore required the smallest amount of damage (see Eqn. (5)). In fact, the damage was not sufficient to initiate a fracture. To maintain the match at late times, it was necessary to increase Rmin with time (reduce cumulative damage). Although we do not have a physical explanation for this phenomenon, it is in agreement with the conventional skin analysis shown in Fig. 9 where the skin decreases from 100 to 65 during the second year. Case 2 had larger thermal stress T and required more damage (pe) to match pf. However, because of the low fracture conductivity, the friction pressure

SPE 95695

COUPLED SIMULATION OF RESERVOIR FLOW, GEOMECHANICS AND FORMATION PLUGGING

BHP inj pressure (psia)

pfric was significant, and the altered stress (i.e., c - T + pe) required was lower than pf. This resulted in a small fracture of about 9 ft. Case 3 required much more damage as the friction pressure was small and the altered stress needed to be higher than in Case 2 and close to the injection pressure. The effect of higher damage was to accelerate fracture propagation; the length at the end was 366 ft.
Case 1 - median aL 2 - high aL, high conductivity 3 - high aL, low conductivity (1/ft)
n

Case 2 Forecast, infinite reservoir, BHP for various rates, DT=27 deg F
3400

3200

Q=100,000, BH inj pressure Q=125,000, BH inj pressure Q=150,000, BH inj pressure

3000

2800

n 1 1 1

Rmin 0.1 0.035 0.003

2600

0.0005 0.002 0.2

2400

2200

Table 1 Damage parameters for the 3 matches


2000

BHP inj pressure (psia)

The damage in the first two cases is localized close to the injector, while in Case 3 it reaches far into the reservoir. In this respect, Case 3 is considered less realistic. Apart from the oscillations due to fracture crossing block boundaries (which could be reduced by still finer gridding), the three matches are of similar quality, yet they produce a very different picture of the process. This demonstrates the danger of using coupled (or any) modeling with too little data. The ambiguity can be however dealt with as discussed below. Predictions for typical future injectors At this preliminary stage, all three scenarios were used to generate forecasts for injection rates up to 150,000 BWPD, and with different T, with the aim to determine the safe surface injection pressure (THP) specifications. At higher rates, all scenarios produced fracturing, but the predicted surface pressures were somewhat different. For Case 1, the damage is low such that even the low fracture transmissibility results in high dimensionless fracture conductivity. The predicted BHP is insensitive to rate as shown on Fig. 11 for the case of T =27 oF. Therefore THP is primarily a function of wellbore friction. For Case 2 there is a significant dependency on rate as well as on fracture length, and the BHP as well as THP continues to increase with time, as shown on Fig. 12 T =27 oF. For Case 3 the pressure also climbs with time. In all cases, injection temperature has a large effect on BHP as shown in Fig. 13 for Case 1.
Case 1 Forecast - Infinite Reservoir, BHP for Various Rates for DT=27 deg F
3400

200

400

600

800

1000

1200

1400

1600

1800

2000

time (days)

Fig. 12 BHP dependence on rate for Case 2, T =27 oF


Case 1 Forecast - Infinite Reservoir, BHP for Various DT, Qw=150,000 stb/d
3400

3200

3000

2800

2600

DT=17 Deg F, BH inj pressure


2400

DT=27 Deg F, BH inj pressure DT=37 Deg F, BH inj pressure

2200

2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000

time (days)

Fig. 13 BHP dependence on T, Case 1, Q=150,000 BWPD Design consequences It is obvious that due to the complexity of the process, as many uncertainties as possible must be eliminated to design high rate re-injection projects. First, data for aL can be easily measured. Second, the damage can be calibrated by matching injection tests below fracture pressure where the fracture aspects do not interfere (like in the GOM example above). Finally, for injection in fracture mode, analysis of step rate/fall-off tests and fracture diagnostics methods can be used to determine fracture dimensions and conductivity by other means, thus leaving only the damage as a matching parameter. The above principles were followed in a subsequent study. As a result, an essentially unique match for Haru 4 was obtained, and the uncertainties of the THP predictions were significantly reduced. As a result, it was possible to recommend ANSI 600 standard for the surface equipment design that will result in significant cost savings for the operator. Conclusions 1) A numerical model solving in a coupled fashion multiphase flow, geomechanics (stress changes), formation plugging, and fracture mechanics has been developed. 2) A simple, flexible model of permeability damage was

3200

BHP inj pressure (psia)

3000

2800

2600

Q=100,000, BH inj pressure


2400

Q=125,000, BH inj pressure Q=150,000, BH inj pressure

2200

2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000

time (days)

Fig. 11 BHP dependence on rate for Case 1, T =27 oF

R.C. BACHMAN, T. HARDING, A. SETTARI AND D.A. WALTERS

SPE79695

formulated and implemented in the reservoir flow part of the system. It is capable of reproducing the injectivity loss observed in Gulf of Mexico and Masila Block wells. 3) The factors controlling injection pressure match are the fracturing, the degree of reservoir permeability damage and thermal stresses, which are dependent on the rock thermal expansion coefficient (T). 4) Multiple interpretations of the field injection pressures are possible, with trade-off between the thermal stress magnitude, poroelastic stress induced by permeability damage, and fracture conductivity. However, if laboratory data on T is available, history match can be used to characterize the plugging mechanics. 5) Accurate prediction of injection pressure is critical as it directly impacts facilities and pipeline design specifications. Forecasting with the calibrated model indicated that 100,000 BWPD injectors should be possible without exceeding ANSI 600 standard for surface equipment design. 6) Due to the cumulative permeability damage, the only feasible method of maintaining injectivity in high rate produced water re-injection is sustained fracturing. Nomenclature A = flow area (m2) aL = linear thermal expansion coefficient (1/deg C) k = permeability (md) k0 = undamaged permeability (md) n = exponent in damage equation pf = fracture propagation pressure (kPa) Q = injection rate (m3/d) Rmin = maximum damage parameter in Eqn. (4) V = cumulative water flow through area A (m3) T pe pfric T c = damage strength parameter in Eqn. (4) = coefficient in Eqn. (1) = thermal stress (kPa) = poroelastic stress (kPa) = friction pressure in the fracture (kPa) = temperature difference (reservoir-inj) (deg C) = undisturbed confining stress on fracture (kPa) = particle concentration (kg/m3)

4. 5. 6.

7.

8.

9.

10. 11. 12. 13.

Acknowledgements The authors wish to thank Nexen Inc. and its partners in the Masila Project, Occidental Petroleum Ltd., Consolidated Contractors Company S.A.L. and the Government of the Republic of Yemen, for the permission to publish this paper. References 1. Sharma, M.M., Pang, Shutong, Wennberg, K.E. and Morgenthaler, L. : Injectivity Decline in Water Injection Wells: An Offshore Gulf of Mexico Case Study, Paper SPE 38180, 1997 SPE European Formation Damage Conference, The Hague, Netherlands, 2-3 June 1997. 2. Martins, J.P., Murray, L.R., Clifford, P.J., McLelland, W.G, Hanna, M.F and Sharp, J.W.: Produced Water ReInjection and Fracturing in Prudhoe Bay, SPE Reservoir Engineering, August 1995, 176-182. 3. Settari, A. and Walters, D.A.: Advances in Coupled

14.

15.

16. 17.

Geomechanical and Reservoir Modeling With Applications to Reservoir Compaction, SPE Journal, Vol. 6, No. 3, Sept. 2001, pp. 334-342. Settari, A. and Mourits, F.M.: A Coupled Reservoir and Geomechanical Simulation System", SPE Journal, September 1998, pp. 219-226. Settari, A.: Reservoir Compaction, Dist. Author Series, J. Pet. Technol., August 2002, pp. 62-69. Heffer, K.J., Last, N.C., Koutsabeloulis, N.C., Chan, H.C.M., Gutierrez, M. and Mukarat, A.: The Influence of Natural Fractures, Faults and Earth Stresses on Reservoir Performance Geomechanical Analysis by Numerical Modelling, North Sea Oil and Gas Reservoirs III, pp. 201-211, Proceedings 3rd Int. Conf. On North Sea Oil and Gas Reservoirs, Trondheim, Kluwer Acad. Publ., 1994. van der Hoek, P.J.: A Simple and Accurate Description of Non-linear Fluid Leak-off in High Permeability Fracturing, paper SPE 63239, SPE Annual Tech. Conf., Dallas, TX. 1-4 Oct., 2000. Gheissary, G., Fokker, P.A., Egberts, P.J.P., Floris, F.J.T., Sommerauer, G. and Kenter, C.J.: Simulation of Fractures Induced by Produced Water Re-Injection in a Multi-Layer Reservoir, paper SPE 54735, SPE European Formation Damage Conf., The Hague, Netherlands, 31 May-1 June, 1999. Saripalli, K.P., Gadde, P.B., Bryant, S.L. and Sharma, M.M.: Role of Fracture Face and Formation Plugging in Injection Well Fracturing and Injectivity Decline, paper SPE 52731,SPE/EPA E&P Environm. Conf., Austin, TX. 28 Feb-3 March, 1999. Settari, A.: Simulation of the Hydraulic Fracturing Processes, SPEJ, (Dec. 1980). pp. 487-500. Nghiem, L.X., Forsyth, Jr. P.A., and Behie, A.: A Fully Implicit Hydraulic Fracture Model, J. Pet. Tech., 1984, pp. 1191. Settari, A., Puchyr, P.J. and Bachman, R.C.: Partially Decoupled Modelling of Hydraulic Fracturing Processes', SPE PE, February 1990, pp.37-44. Settari, A. and Warren, G.M. : Simulation and Field Analysis of Waterflood Induced Fracturing, Paper SPE/ISRM 28081, Proceedings, SPE/ISRM Meeting "Rock Mechanics in Petroleum Engineering", Delft, Aug.29-31, Balkema Publ., 1994, pp. 435-445. Settari, A., Warren, G.M. Jacquemont, , J., Bieniawski, P. and Dussaud, M.: Brine Disposal into a Tight Stress Sensitive Formation at Fracturing Conditions: Design and Field Experience, SPE Reservoir Eval. & Eng., Vol 2, No. 2, April 1999, pp.186-195. Settari, A., Sullivan, R.B., and Bachman, R.C.: The Modeling of the Effect of Water Blockage and Geomechanics in Waterfracs, Paper SPE 77600, presented at the Annual Techn. Conf. of SPE, San Antonio, TX., Sept. 29 Oct. 2, 2002. Pang, S. and Sharma, M.M.: A Model for Predicting Injectivity Decline in Water Injection Wells, SPE Form. Eval., pp. 194-201, Sept., 1997. Wennberg, K. E.: Particle Retention in Porous Media: Applications to Water Injectivity Decline, PhD Thesis, Dept. of Petroleum Engineering and Applied Geophysics, The Norwegian University of Science and Technology,

SPE 95695

COUPLED SIMULATION OF RESERVOIR FLOW, GEOMECHANICS AND FORMATION PLUGGING

Trondheim, Feb. 1998. 18. Soo, H. and Radke, C.J.: A Filtration model for Flow of Dilute, Stable Emulsions in Porous Media 1. Theory, Chem. Eng. Sci., Vol. 41, No. 2, 1986, 261-272. 19. Sharma, M. M. and Yortsos, Y.C.: Transport of Particulate Suspensions in Porous Media: Model Formulation, AIChE J., Vol. 33, Oct. 1987, 1636-1643. 20. Bedrikovetsky, P., Marchesin, D., Shecaira, F.S., Souza, Antonio Luiz S., Milanez, Paulo V. and Rezende, Emerson: Injectivity Decline Caused by Injection of Sea/Produced Water: Applications to Waterflood Management, Paper IBP 43300, Rio Oil and Gas Conference, Rio de Janeiro, Oct. 16-19, 2000. 21. Harding, T G., Smith, K.H., Al-Hakimi, E., Al-Seyani, A., Wilkie, D, and Willson, N.D.: Produced Water Management: Masila Block Yemen, presented at the 2nd International Yemen Oil & Gas Conference, Sanaa, Republic of Yemen, 24-25 June 2002. 22. Harding, T.G., Smith, K.H., and Norris, B.: Horizontal Water Disposal Well Performance in a High Porosity and Permeability Reservoir, paper SPE 79007, presented at the International Thermal Operations and Heavy Oil Symposium and International Horizontal Well Technology Conference, Calgary, 4-7 November 2002.

Das könnte Ihnen auch gefallen