You are on page 1of 9

CSIRO PUBLISHING

www.publish.csiro.au/journals/eg

Exploration Geophysics, 2010, 41, 146154

A comparison of rock physics models for uid substitution in carbonate rocks


Ali Misaghi1,3,5 Sajjad Negahban2 Martin Landr3 Abdolrahim Javaherian4
1 2

Dana Geophysics Company, Golestan Street, Iran Zamin Avenue, Shahrak-e-Qods, Tehran 1465865187, Iran. Amirkabir University of Technology, No. 242, Haz Street, Tehran, P.O.B. 15875-4413, Iran. 3 Norwegian University of Science and Technology, S.P. Andersens vei 15A, 7491 Trondheim, Norway. 4 Institute of Geophysics, University of Tehran, North Karegar Avenue, Tehran 1435944411, Iran. 5 Corresponding author. Email: ali.misaghi@gmail.com

Abstract. Rock physics models play a crucial role in seismic reservoir characterisation studies. The optimal rock physics model for a sandstone reservoir might be signicantly different from that of a carbonate reservoir. There are several theories that compare the elastic properties of dry and saturated rocks. These models have mainly been explained by poroelastic theories or effective medium theories. The Gassmanns model which is commonly used in petroleum rock physics is suitable for rocks with spherical and interconnected pores at low frequencies. These assumptions do not necessarily meet the conditions of carbonate rocks. In this work, two additional models, the differential effective medium (DEM) model and the self-consistent (SC) model have been examined for several carbonate samples. Ultrasonic 30 carbonate and 5 sandstone core samples from an oileld in south-west Iran were measured in the laboratory. The results show that the DEM model gives the best compatibility with the dense and low porous carbonate samples. These results are conrmed by well log data from the same area. Key words: carbonate reservoirs, differential effective medium model, Gassmanns model, rock physics, self-consistent model.

Introduction Today, seismic reservoir monitoring is an established tool for monitoring of enhanced oil recovery processes and production management. It is evident, especially for sandstone reservoirs, that seismic methods are applicable for the entire life cycle of a reservoir, instead of only for the exploration phase. In seismic reservoir monitoring, four typical types of reservoir changes are studied: saturation changes, pressure changes, geo-mechanical changes (both within and outside the reservoir) and temperature changes. A major question is how these physical changes can be related to changes in seismic properties of the medium. The answer is often given by rock physics models, which can give more insight into these relationships. There are several important theories in rock physics, of which Gassmanns model probably is the most well known. The Biot and Gassmann theories are used to predict how P- and S-wave velocities are changed as saturation changes. In these theories, the underlying assumptions are discussed by White (1983), Bourbi et al. (1987), and Mavko et al. (1998). Gassmanns theory applies to the low-frequency case (Gassmann, 1951) and it is identical to the low-frequency limit of Biots theory (Biot, 1956a). For the high-frequency range, Biot (1956b) derived theoretical expressions for predicting P- and S-wave velocities of saturated rocks in terms of dry rocks. Biots theory predicts two coupled P-waves and one S-wave for the case of an isotropic, liquid-saturated porous medium. Wyllie et al. (1963) indicated that only those results should be considered reliable that show no increase in S-wave velocity when the dryrock specimens are saturated with liquids. King (1966) reported cases of ultrasonic S-wave velocities for brine- and kerosenesaturated rocks at lower conning stresses that are higher than dry rock S-wave velocities. Mavko et al. (1998) suggested that Biots
ASEG 2010

theory for high frequencies is valid in the case of very highpermeable porous materials, such as ocean sediments. Plona (1980) demonstrated the existence of Biots slow P-wave in a porous, high-permeable material made of sintered glass spheres (see also Bourbi et al., 1987; King, 2005). Experimental data on carbonate rocks have not been as thoroughly studied as silisi-clastic sedimentary rocks. Among others Domenico (1984), Rafavich et al. (1984), Anselmetti and Eberli (1993), Assefa et al. (2003), Prasad and Nur (2003), Baechle et al. (2005) and Adam et al. (2005) described the ultrasonic velocity response of carbonate rocks to porosity, permeability, texture, uids and pressure. However, there are relatively few studies on carbonate rock physics models. There are several rock physics models for estimation of seismic velocity versus saturation. Since the Gassmann model is based on some specic assumptions, the concordance of rocks and uid properties with these assumptions should be evaluated before its application. In addition to Gassmanns method, there are some other theories, which can be used in the estimation of the bulk and shear moduli and ultimately P- and S-wave velocities. Effective medium theories, self-consistent (SC) theories and differential effective medium (DEM) theories are some of the models which can be used in rock physics. In this study, rst, Gassmanns theory and its failure points in carbonate rocks will be reviewed and second, alternative rock physics models for carbonate rocks will be investigated. Rock physics models Gassmanns model The most widely-used theory for uid substitution is the lowfrequency Gassmanns theory. Gassmanns equation gives
10.1071/EG09035 0812-3985/10/020146

Comparison of rock physics models for uid substitution in carbonate rocks

Exploration Geophysics

147

a relationship between saturated bulk modulus, porosity, bulk modulus of rock frame, bulk modulus of minerals of rock matrix and the bulk modulus of pore uids (Mavko et al., 1998):  2 k 1 kdry m 1 k sat k dry  1 k dry ; kfl km k 2
m

where, ksat is the saturated bulk modulus, kdry is the bulk modulus of rock frame, km is the bulk modulus of minerals of the rock matrix, k is the bulk modulus of the pore uids and  is the porosity. Gassmanns equation is based on several assumptions, which must be taken into account in any application (Wang, 2001): (1) rock (matrix and frame) must be macroscopically homogeneous; (2) all pores must be interconnected; (3) pores are lled with a frictionless uid; (4) the rock-uid system must be closed (undrained); (5) there should be no interaction between uid and the matrix in a way that could soften or harden the frame. The rst assumption implies that the wavelength must be greater than the pore and grain sizes. The second assumption indicates that the porosity and the permeability must be high and there should be no isolated or poorly connected pores. Assumptions 2 and 3 explain why the log and laboratory velocity data often are higher compared to Gassmanns predictions. For some frequencies, a relative movement between the uid and the matrix will occur, and this might lead to dispersive waves. Relative uid-matrix movements are generally more prominent for some special (resonance like) frequencies, and might create large differences between bulk and shear moduli of uid and matrix. With these assumptions and equation 1, saturated bulk modulus, ksat, can be estimated. By knowing ksat, P- and S-wave velocities can be predicted by equations 2 and 3, (Mavko et al., 1998): s k sat 4=3m 2 ; Vp rB r m Vs 3 rB ; where, m and rB are the shear modulus and bulk density, respectively. Now, we discuss some of the main characteristics of carbonate rocks which do not meet Gassmanns assumptions. Heterogeneity and pore types Carbonate rocks usually show large heterogeneity, vugs and mouldic structures that have comparable length with the ultrasonic wavelength. Due to the presence of different inclusions and elements with different densities and pores with large sizes in the carbonate rocks, signicant scattering of the ultrasonic waves may occur in these rocks (Adam et al., 2006). New studies have shown that carbonate rocks with spherical pores are less sensitive to pressure changes for seismic frequencies, and that these are in good accordance with Gassmanns equation. Despite this, carbonate rocks with soft and fracture-type pores which are sensitive to stress do not follow the Gassmanns theory (Adam et al., 2006). Variation of shear moduli Constancy of dry and saturated shear moduli are not part of Gassmanns assumptions, however, these can be derived from Gassmanns equation (Berryman, 1999). In Gassmanns equation only the bulk modulus is changing by saturating the rock, but some studies have shown that the shear modulus may not be constant and may vary under some conditions (Adam et al., 2005).

Some laboratory studies have also validated that shear modulus is varying in dry and saturated rocks (Vo-Thanh, 1995; Assefa et al., 2003; Baechle et al., 2005; Sharma et al., 2006). In rocks with intergranular and intercrystalline porosities, m decreases as saturation increases, while in rocks with mouldic porosity or micro porosity, saturation has less effect on velocities (Baechle et al., 2005). Therefore, investigation of constancy of shear modulus can be a way to check the accuracy and applicability of Gassmanns equation for a specic rock. Pressure, pore connectivity, dispersion and clay content For low effective pressure, since the fractures and cracks are open, the differences between velocities estimated by Gassmanns equation and measured values are increased for a saturated rock. High effective pressure will close the soft fractures and pores. Consequently Gassmanns equation may estimate the results with less error (Adam et al., 2005). Gassmanns assumptions state that the pores must be connected to each other. Since in most carbonate rocks many isolated pores exist, Gassmanns assumptions are not necessarily valid (Adam et al., 2005). Usually higher frequencies show larger moduli (Adam et al., 2006) and dispersion of moduli is higher in saturated rocks than in dry rocks (Adam et al., 2005). The shear modulus at sonic frequencies is reduced when the rock changes from dry condition to saturated condition, while for ultrasonic frequencies it acts in reverse and the shear modulus increases (Adam et al., 2005; Sharma et al., 2006). Presence of clay in carbonate rocks overestimates the bulk modulus by Gassmanns equation. One reason for this can be the effect of uids on clay structure, which makes them weaker (Adam et al., 2006). On the other hand, clays absorb and contain water and even in drained stages (to make the rocks dry), clay keeps some water and this may change the results (Japsen et al., 2002). Viscosity, pH, compressibility and polarity of uids Fluids must have no viscosity based on Gassmanns assumptions. This assumption is usually broken since most of the reservoir uids are viscous. In addition to viscosity, pH of the uid should be considered, since this might also affect the interaction between the uid and the rock and ultimately change the estimated shear and bulk moduli. More compressible uids make the differences between dry and saturated samples smaller and show more compatibility with Gassmanns model. Changes in shear modulus are smaller in non-polar uids and the measured results are in better accordance with Gassmanns predictions (Adam et al., 2006). Some of the failure points for the Gassmann model in carbonate rocks have been reviewed and the necessity of a new model with better match with the characteristics of carbonate rocks seems crucial. Effective medium models The effective medium theory assumes that separated pores and cracks may or may not be connected to each other (Berryman et al., 2002). Self-consistent and DEM theories are part of the effective medium theories. In this case, detailed knowledge regarding pore shape and characteristics of microstructures is necessary. These two theories have two major advantages: ease of calculation and exibility of application. The SC model The SC model was introduced by Kerner (1956). Later, Hill (1965) and Budiansky (1965, 1970) computed the elastic properties of a two-phase medium. This method is based upon the following idea: a single inclusion representing one of the

148

Exploration Geophysics

A. Misaghi et al.

components is embedded within a large surrounding matrix whose elastic properties are those of the effective medium. Wus SC modulus estimation for two-phase composites may be expressed as equations 4 and 5 (m = matrix, I = inclusion), Wu (1966). Berryman (1980a, 1980b, 1995) presented a more general form of the self-consistent approximation for N-phase composites in equations 6 and 7 (Mavko et al., 1998). K *SC Km xi Ki Km P*i m*SC mm xi mi mm Q*i
N X i 1 N X i 1

4 5 6 7

xi Ki K *SC P*i 0 xi mi m*SC Q*i 0

Where i refers to the ithmaterial, xi is its volume fraction, P and Q are geometric factors, superscript *i on P and Q indicates that the factors are for an inclusion of material i in a background medium with SC effective moduli K*SC and, m*SC, where K and m are the bulk and shear moduli of an uncracked medium, respectively. DEM model Differential effective medium theory takes the point of view that a composite material may be constructed by making innitesimal changes in an already existing composite. We suppose that there are only two constituents in the medium, whose volume fractions are x = v(1) and y = v(2) = 1 x. Furthermore, we assume that the value of the effective bulk modulus K*DEM(y) at one value of y is known. Treating K*DEM(y) as the composite host medium and K*DEM(y + dy) as the effective constant after a small proportion dy/(1 y) of the composite host has been replaced by spherical inclusions of type (2), therefore (Berryman and Berge, 1993): K *DEM y dy K *DEM y dy k 2 K *DEM y : 8 1 y k 2 4 K *DEM y dy 4 3 m*DEM y 3 m*DEM y Since the composite host contains the volume fraction x of type (1) and y of type (2), on average a fraction dy/(1 y) of the composite host must be replaced by type (2) in order to change the overall fraction of type (2) to y + dy. Taking the limit dy ! 0 gives the following results: 1 y 1 y d k *y k 2 k *P*2 y; dy d m*y m2 m*Q*2 y: dy 9 10

carbonate rocks is independent of mineralogy. Velocity variation in carbonate rocks is a complex function of several parameters. These parameters can be categorised in two main groups: internal and external parameters. Internal parameters like porosity, pore type, composition and grain sizes are related to the lithology of the rock and physical properties of fabric of the rock. Induced pressures (by overburden or tectonic) and physical characteristics of the passing waves are some examples of external parameters. In addition to porosity, type and shape of the pores also impact the seismic velocity in sedimentary rocks. Basically seismic velocities are the amount of deformability (compressibility and rigidity) of rocks, which are dependent on shape and number of pores. A at thin pore deforms easily and then the seismic velocity is low in the rocks with this type of pores. In contrast, round and circular pores are resistant to deformation, therefore seismic velocity is high in rocks with spherical pores, although they may have high porosity (Palaz and Marfurt, 1997). Porosity in carbonate rocks can be classied in several main groups: intergranular and intercrystalline; mouldic and interparticle; vugs and channels; and nally fenestral. Seismic velocity is different from one porosity type to another. In the intergranular and intercrystalline porosity group, seismic velocities are low and hence more sensitive to the effective pressure while in rocks with interparticle and mouldic porosities, seismic velocities are relatively high and are not so sensitive to the effective pressure, since these types of pores are very resistant against deformation. Rocks with vugs usually have a rigid framework and P- and S-wave velocities are high in these rocks. Also due to the rigid framework and low ratio of internal surfaces to the porosity they are not very sensitive to pressure. Rocks with channel shape pores are easily deformed, especially when dissolution of fractures and pores make these channels. Therefore, these rocks show low seismic velocities in certain porosities by increasing the pressure; P- and S-wave velocities may increase initially since some of the pores may collapse. Seismic velocities in rocks with fenestral pores behave similarly to the rocks with intercrystalline pores, and in certain porosities and pressures show low P- and S-wave velocities, but by increase in pressure, due to deformation of pores, the velocities will increase. Laboratory measurements In this study, some data from the Sarvak Formation, a carbonate reservoir in south-west Iran, were used. Thirty carbonate core samples and ve sandstone samples were tested at different saturations and at constant pressure and temperature conditions (4400 psi and 95C) for P-wave velocity measurements using ultrasonic waves. The measured parameters on the core samples have been listed in Table 1. In a typical section, the Sarvak Formation (Cenomanian) includes three units with a total thickness of 822 m. The Sarvak Formation consists of carbonate sediments deposited in a shallow marine environment. The average porosity of the Sarvak Formation is ~18% and the average permeability is 45 mD in the reservoir zones. To be able to compare the efciency of different rock physics models, an attempt has been made to keep the conditions of the experiments similar to each other, and also to use similar constant parameters in the formulations for different rock physics models. The origin of the parameters that were used in the different rock physics models are summarised in Table 2. Comparing measured data with the Gassmanns model By measuring P- and S-wave velocities and knowing rock and uid properties, saturated bulk modulus can be estimated using

With initial conditions k*(0) = k1,*m(0) = m1, k and m1 are bulk and shear moduli of the host material (phase 1), bulk and shear moduli of the incrementally added inclusion (phase 2), and y is concentration of phase 2. For uid inclusions and voids, y equals the porosity. The terms P and Q are geometric factors (Mavko et al., 1998). Seismic characteristics of carbonate rocks Seismic velocity in carbonate rocks is controlled by two main factors: lithology of sediments and sedimentation processes like: cementation and dissolution. Carbonate rocks, in comparison with other sedimentary rocks, show less compositional variations. Calcite, dolomite and aragonite are the main (more than 95%) minerals in carbonate rocks. These minerals have similar physical properties, therefore, the wide range of velocity variations in

Comparison of rock physics models for uid substitution in carbonate rocks

Exploration Geophysics

149

Table 1. The measured parameters on the core samples. Sample Lithology Depth (m) Bulk density (g/cc) 2.09 2.27 2.06 2.40 2.53 2.08 2.08 2.17 2.10 2.14 2.02 1.93 2.50 2.57 2.29 2.17 2.24 2.20 2.05 2.23 1.98 1.92 2.01 2.60 2.56 2.37 2.22 2.42 2.59 2.37 2.20 2.59 2.61 2.41 2.52 Grain density (g/cc) 2.70 2.66 2.70 2.64 2.76 2.65 2.68 2.66 2.65 2.66 2.70 2.71 2.67 2.65 2.69 2.70 2.70 2.70 2.71 2.70 2.68 2.72 2.74 2.69 2.69 2.70 2.69 2.69 2.69 2.70 2.70 2.69 2.70 2.68 2.67 Porosity (%) Ave. horizontal permeability (mD) 2.86 0.81 6.32 0.23 0.17 490.75 362.74 806.29 2208.35 1364.71 74.40 73.04 0.10 0.03 1.09 1.89 23.96 6.13 59.78 7.67 179.15 95.93 228.15 2.38 0.35 2.24 70.54 0.46 0.08 3.47 10.12 0.66 0.12 1.07 0.51 Effective pressure (psi) 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 4400 Temp. (C) Dry Vp (km/s) Saturated Vp (km/s)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35

Carbonate Carbonate Carbonate Carbonate Carbonate Sandstone Sandstone Sandstone Sandstone Sandstone Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate Carbonate

2881.1 2892.5 2894.2 3783.1 3786.6 3839.4 3839.9 3844.3 3844.5 3845.8 2765.2 2770.0 2870.0 2878.0 2882.4 2781.4 2788.0 2789.3 2802.6 2808.2 2811.5 2817.4 2819.1 2832.3 2842.3 2845.9 2849.7 2638.4 2648.6 2654.5 2656.0 2676.7 2660.1 2684.3 2694.1

22.43 14.51 23.92 8.86 8.19 21.47 22.38 18.34 20.96 19.58 25.25 28.62 6.61 2.89 14.89 19.49 16.97 18.74 24.44 17.41 26.30 29.45 26.82 3.29 4.76 12.34 17.21 9.94 3.85 12.41 18.42 3.99 3.26 10.24 5.53

95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95 95

3.54 4.11 3.35 4.58 4.42 3.76 3.75 3.87 3.84 3.69 3.44 3.94 4.70 5.25 3.76 4.13 4.16 3.89 3.51 3.98 3.79 3.48 3.49 5.59 5.22 4.66 4.69 4.57 4.85 4.32 4.42 5.47 5.15 4.67 5.02

3.85 4.53 3.68 4.61 4.93 3.86 3.71 3.97 3.92 3.81 3.88 4.19 4.81 5.47 4.28 4.26 4.35 4.17 3.73 4.37 4.09 3.68 3.81 6.02 5.48 4.79 4.81 4.82 5.44 4.69 4.53 5.48 5.73 4.94 5.27

Table 2. Origin of parameters used for different models. Model Km mm Kmineral mmineral K F Kdry mdry vp sat, vs sat, vp dry and vs dry Ksat msat = mdry rb dry rb sat F (volumetric fraction of mineral) vp sat model vs sat model P*i, Q*i Variable: Gassmann Smith et al. (2003) Equations 9, 10, 11 Carmichael (1989) Carmichael (1989) Laboratory measurements Laboratory measurements Smith et al. (2003) Equation 2 Smith et al. (2003) Equation 3 Laboratory measurements Smith et al. (2003) Equation 1 Berryman (1999) Laboratory measurements Laboratory measurements Laboratory measurements Smith et al. (2003) Equation 13 Smith et al. (2003) Equation 14 DEM Smith et al. (2003) Equations 9, 10, 11 Smith et al. (2003) Equations 9, 10, 11 Carmichael (1989) Carmichael (1989) Laboratory measurements Laboratory measurements Mavko et al. (1998) Mavko et al. (1998) Laboratory measurements Laboratory measurements Smith et al. (2003) Equation 13 Smith et al. (2003) Equation 14 Table 4.9.1 Mavko et al. (1998) SC Smith et al. (2003) Equations 9, 10, 11 Smith et al. (2003) Equations 9, 10, 11 Carmichael (1989) Carmichael (1989) Laboratory measurements Laboratory measurements

Mavko et al. (1998) Laboratory measurements Laboratory measurements Smith et al. (2003) Equation 13 Smith et al. (2003) Equation 14 Table 4.8.1 Mavko et al. (1998)

150

Exploration Geophysics

A. Misaghi et al.

Gassmanns equation 1. Based on Gassmanns assumptions, shear modulus is supposed to be equal to the dry case. Now, using equations 2, P-wave velocities can be calculated for the saturated case. To see how well the Gassmann model ts the studied samples, the measured P-wave velocity in the laboratory for water saturated samples were compared with the results from Gassmanns equation (Figure 1). In the presented graphs, the equation of the best tted curve to the data, R2, root mean squared error (RMSE), sum of squares due to error (SSE) and sum of squared error (S) have been calculated. Figure 1 shows that, except for one sample, Gassmanns model predicts lower P-wave velocities than those measured. The mismatch between Gassmanns assumptions and the characteristics of the samples and experiment conditions (like measurements at high frequencies, variation of shear modulus in dry and saturated conditions and improper pore connections) can be some reasons for the discrepancies. As expected Gassmanns model gives better correspondence with measurements for the sandstone samples, even though these are not completely clean. Since the pores in these carbonate samples have spherical shape, the data from Gassmanns model are not so dispersed in this study. Nevertheless, the differences between measured values and predicted values from Gassmanns model are signicant. The relationships between porosity and P-wave velocity; porosity and bulk modulus for measured and estimated (Gassmann) values are shown in Figure 2. The apparent similarity in shape of the graphs (not the values) is interesting and spherical pore shapes can be the cause for these similarities. However, for all graphs, Gassmanns model shows lower values.

Vp sat Gassmann (km/s)

Carbonate sample Best fit Sandstone sample

y = 0.074*x7652 R 2 = 0.95 RMSE = 0.1574 SSE = 0.6938

2 2

Vp sat measurement (km/s)


Fig. 1. A comparison between P-wave velocity calculated by Gassmanns model and measured in the laboratory for 32 carbonate samples (red circles) and 5 sandstone samples (blue squares) in an oileld in south-west Iran. Green dashed line is the best t passing through the data with its equation. Black continuous line is the ideal line when the velocity calculated by the model and measured data in the laboratory are the same. RMSE, root mean squared error; SSE, sum of squares due to error.

Comparing measured data with the SC model The SC theory allows a composite to become disconnected at some nite porosity; therefore the SC theory is expected to provide good elastic property estimates for granular sedimentary rocks and other materials that become unconsolidated at high

Bulk modulus sat measurement (GPa)

Vp sat measurement (km/s)

y = 0.002*x 2 0.13*x + 6 R 2 = 0.8917 RMSE = 0.2218 SSE = 1.3288

70 60 50 40 30 20 10 0 0 5 10 15 20 25 30 35
y = 0.064*x 2 3.2*x + 64 R 2 = 0.8942 RMSE = 4.151 SSE = 465.20 Carbonate sample Best fit Sandstone sample

Carbonate sample Best fit Sandstone sample

10

15

20

25

30

35

Porosity (%)
7

Porosity (%)

Vp sat Gassmann (km/s)

y = 0.017*x 2 0.13*x + 5.6 R 2 = 0.8806 RMSE = 0.2561 SSE = 1.7707

Bulk modulus sat Gassmann (GPa)

70 60 50 40 30 20 10 0 0 5 10 15 20 25 30 35
y = 0.047*x 2 2.6*x + 51 R 2 = 0.8637 RMSE = 4.246 SSE = 486.66 Carbonate sample Best fit Sandstone sample

Carbonate sample Best fit Sandstone sample

3 0 5 10 15 20 25 30 35

Porosity (%)

Porosity (%)

(a)

(b)

Fig. 2. Analogy between Gassmanns model and the measured data in the laboratory for 32 carbonate samples (red circles) and 5 sandstone samples (blue squares) in an oileld in south-west Iran: (a) analogy for P-velocity versus porosity and (b) analogy for bulk modulus versus porosity, green dashed line is the best t passing through the data. RMSE, root mean squared error; SSE, sum of squares due to error.

Comparison of rock physics models for uid substitution in carbonate rocks

Exploration Geophysics

151

porosities. For samples with microstructure like fuse glass beads, the SC model is expected to show good estimates. Because assumptions of this model are more compatible with microstructures of sandstone, this model gives better results in these rocks. But for carbonate samples that have continuous matrix and discontinuous inclusions, the SC model seems unsuitable. This is in agreement with the measurements shown in Figure 3, where a signicant discrepancy between the SC model and the ultrasonic measurements is observed. This discrepancy is probably due to the SC assumption which is not compatible with typical carbonate rock microstructure, because matrix of this sample is continuous and inclusions are discontinuous. Furthermore, these samples are not granular. For sandstone samples in this study, the results show that the SC model does not give suitable estimations, which might be caused by the relatively high shale content of the samples. Figures 3 and 4 show that the SC model overestimates the velocities and modulus for these samples. However, contrary to the Gassmanns model there is no similarity between analogue charts.

Comparing measured data with the DEM model The DEM model is expected to give better results for the samples in our study because this model was developed for glass foam. The microstructure of glass foam is similar to that of carbonates so that the basic assumptions underlying the DEM model better match our data. One of the main assumptions of this model is that the particles should be discontinuous and the matrix continuous. This means that the cracks and pores are isolated, which is the case for most carbonate rocks. Figure 5 shows that the results from the DEM model are well correlated with the measured data for the carbonate samples. However, for the sandstone samples there is less correlation, and we think this is due to differences in microstructures partly violating the assumptions of the model. Analysis of the errors of the three different models reveals that the DEM model is best choice for carbonate rocks. If the error S is calculated as the sum of squared differences of the measured velocity of compressional waves in the laboratory and calculated using the models, S equals 6.13, 41.02 and 4.07 for Gassmanns, SC and DEM models, respectively. Figure 6 shows another sum of the squared errors (S). Here, S is the sum of the squared
7
Carbonate sample

7.0 6.5

Vp sat SC (km/s)

6.0

Best fit Sandstone sample

5.5 5.0 4.5 4.0 3.5 3.5 4.0 4.5 5.0 5.5

Carbonate sample Best fit Sandstone y = 0.808*x+2.01 2 R = 0.7802 RMSE = 0.2839 SSE = 2.257

Vp sat DEM (km/s)

4
y = 1.29*x 1.29 2 R = 0.88 RMSE = 0.3232 SSE = 2.92

6.0

6.5

7.0

2 2 3 4 5 6 7

Vp sat measurement (km/s)


Fig. 3. A comparison between P-wave velocity calculated by self-consistent (SC) model and measured in the laboratory for 32 carbonate samples (red circles) and 5 sandstone samples (blue squares) in an oileld in southwest Iran. Green dashed line is the best t passing through the data with its equation. Black continuous line is the ideal line when the velocity calculated by the model and measured data in the laboratory are the same. RMSE, root mean squared error; SSE, sum of squares due to error.

Vp sat measurement (km/s)


Fig. 5. A comparison between P-wave velocity calculated by differential effective medium (DEM) model and measured in the laboratory for 32 carbonate samples (red circles) and 5 sandstone samples (blue squares) in an oileld in south-west Iran. Green dashed line is the best t passing through the data with its equation. Black continuous line is the ideal line when the velocity calculated by the model and measured data in the laboratory are the same. This gure shows DEM is a suitable model for the carbonate samples. RMSE, root mean squared error; SSE, sum of squares due to error.
7
y = 0.0019*x 10.6*x+75 2 R = 0.99 RMSE = 1.3 SSE = 45.64
2

Bulk modulus sat SC (GPa)

80

60

40

Vp sat models (km/s)

20

Carbonate sample Best fit Sandstone sample

4
Gassmann S = 6.13 DEM S = 4.075

0 0

10

15

20

25

30

35

3
Best result S = 0 SC S = 41.02

Porosity (%)
Fig. 4. Bulk modulus calculated by the self-consistent (SC) model and measured in the laboratory for 32 carbonate samples (red circles) and 5 sandstone samples (blue squares) in an oileld at south-west Iran. Green dashed line is the best t passing through the data with its equation. Black continuous line is the ideal line when the velocity calculated by the model and measured in the laboratory are the same. RMSE, root mean squared error; SSE, sum of squares due to error.
2 2 3 4 5

Vp sat measurement (km/s)


Fig. 6. S-values for the best t lines for each model (dashed lines) and the ideal line when the velocity calculated by the model and measured in the laboratory are the equal (black continuous line).

152

Exploration Geophysics

A. Misaghi et al.

differences between the best tted lines through each model data (measured velocity of compressional waves in laboratory and calculated using the models) from the best result line (when measured velocity of compressional waves in laboratory and calculated using the models are equal). These two different S values shows that the DEM model has the minimum S, and thus it can be interpreted as the best model for the carbonate rocks. Comparison between velocity models and log data After comparing the different models using laboratory core samples, the models are compared to log data from the same well as the core samples were taken. Figure 7 shows that the DEM model gives the best estimation while Gassmann and SC models underestimate and overestimate the values, respectively. The accuracy of DEM estimations varies along the log. There is a

good correlation between DEM data and the sonic data in some parts, while other parts show reduced correlation. Investigation of these differences is important and provides useful information. For example a comparison of the composite logs with the estimated velocity logs reveals that the accuracy of the models decreases with increasing shale content. Shales change the matrix of the rocks and thus the physical properties as shales have widely varying shear and bulk moduli which may explain the observed discrepancies. To build an accurate model, the volume of shale and related shear and bulk moduli must be estimated precisely. Presence of oil also reduces the accuracy of the model; this may be due to variations in bulk modulus of oil along the well or inaccurate saturation logs. Presence of oil in shaly layers will accumulate the errors and the models will be less accurate in such areas. Therefore, the logs must be accurately corrected for shale content and correct relative saturation logs must be used in the

LITHOLOGY

GR

PHI

Sw

RHOB

Velocity (k m/s)

Fig. 7. A composite log with estimated velocities using different models. In the velocity column, the red graph is the sonic log; the green graph is the calculated velocity using Gassmanns model, brown for the self-consistent model and blue for the differential effective medium model.

Comparison of rock physics models for uid substitution in carbonate rocks

Exploration Geophysics

153

calculations. Despite the mentioned pitfalls, the DEM model gives the best estimates for carbonate rocks compared to the other models considered in this paper. In Figure 7 in the velocity column, the red graph is the sonic log; the blue graph is the calculated velocity using the DEM model, the green curve corresponds to Gassmanns model and the dark red curve represents the SC model. The blue and red graphs corresponding to DEM and sonic logs, respectively, show the best correlation. Conclusions No model can be recommended as the best model for a given rock or reservoir formation in advance. Microstructure is one of the main factors that control the elastic properties of the rocks, so before any choice of model, the rock microstructure must be studied in detail. Gassmanns model has several underlying assumptions that restrict the models universal use. Gassmann gives a good estimation in rocks with connected pores at low frequencies. By comparing P-wave velocities measured on carbonate core samples for an Iranian oileld with Gassmanns estimated velocities, we have found that Gassmanns estimations are systematically lower than the ultrasonic measurements. In this study, the rock samples have mouldic porosity and due to the spherical shape of the pores Gassmanns model was expected to give good estimates. However, since there is a low connectivity between the pores and also because we measure ultrasonic frequencies in the laboratory, there are discrepancies between modelled and measured results. One point of interest in this comparison is the analogy in shape (not absolute values) of the graphs derived from Gassmanns model and measured data. Selfconsistent assumptions are compatible with fuse glass beads. This model allows composites to be discontinuous in nite porosity. Therefore, this model is suitable for granular rocks and materials that are highly porous and unconsolidated. Carbonate rocks do not satisfy these assumptions as they usually have a continuous matrix; hence the SC model is not suited to model carbonates. The DEM model gives good estimation in microstructures which are similar to glass foam. In this study, the main assumptions are: continuous matrix and discontinuous inclusions, this means pores and cracks have to be isolated. The microstructures of our carbonate samples are compatible with the DEM assumptions. This study showed that indeed the DEM model better matched our data than the other models. The DEM model was ~30 percent more accurate than Gassmanns model. Modelling of uid substitution in porous media is an important step in most reservoir characterisation studies. The results of this study can be used in feasibility studies for time-lapse seismic projects and also in forward modelling workows when we are dealing with uid substitution and saturation changes. Acknowledgement
The authors acknowledge the Research Institute of Petroleum Industry of Iran (RIPI) for permission to use the data. Martin Landr thanks the Norwegian Research Council for nancial support to the ROSE project at NTNU.

References
Adam, L., Batzle, M., and Brevik, I., 2005, Gassmanns uid substitution paradox on carbonates: seismic and ultrasonic frequencies: Presented at the 75th Annual International Meeting, SEG, 24, 15211524. Adam, L., Batzle, M., and Brevik, I., 2006, Gassmanns uid substitution and shear modulus variability in carbonates at laboratory seismic and ultrasonic frequencies: Geophysics, 71, F173F183. doi:10.1190/ 1.2358494

Anselmetti, F. S., and Eberli, G. P., 1993, Controls on sonic velocity in carbonates: Pure and Applied Geophysics, 141, 287. doi:10.1007/ BF00998333 Assefa, S., McCann, C., and Sothcott, J., 2003, Velocity of compressional and shear waves in limestones: Geophysical Prospecting, 51, 113. doi:10.1046/j.1365-2478.2003.00349.x Baechle, G. T., Weger, R. J., Eberli, G. P., Massaferro, J. L., and Sun, Y.-F., 2005, Changes of shear moduli in carbonate rocks: Implications for Gassmann applicability: Leading Edge, 24, 507510. doi:10.1190/ 1.1926808 Berryman, J. G., 1980a, Long-wavelength propagation in composite elastic media, I. Spherical inclusions: The Journal of the Acoustical Society of America, 68, 18091819. doi:10.1121/1.385171 Berryman, J. G., 1980b, Long-wavelength propagation in composite elastic media, II. Ellipsoidal inclusions: The Journal of the Acoustical Society of America, 68, 18201831. doi:10.1121/1.385172 Berryman, J. G., 1995, Mixture theories for rock properties, In T. J. Ahrens, (Ed.), Rock physics and phase relations, A handbook of physics constants, 205228, American Geophysical Union. Berryman, J. G., 1999, Origin of Gassmanns equations: Geophysics, 64, 16271629. doi:10.1190/1.1444667 Berryman, J. G., and Berge, P. A., 1993, Rock elastic properties: Dependence on microstructure: in homogenization and constitutive modeling for heterogeneous materials, In C. S. Cheng, and J. W. Ju, (Eds), AMSE, New York, 113. Berryman, J. G., Pride, S. R., and Wang, H. F., 2002, A differential scheme for elastic properties of rocks with dry or saturated cracks: Geophysical Journal International, 151, 597611. doi:10.1046/j.1365-246X.2002. 01801.x Biot, M. A., 1956a, Theory of propagation of elastic waves in a uid-saturated porous solid, I: Low frequency range: The Journal of the Acoustical Society of America, 28, 168178. doi:10.1121/1.1908239 Biot, M. A., 1956b, Theory of propagation of elastic waves in a uid-saturated porous solid, II: Higher frequency range: The Journal of the Acoustical Society of America, 28, 179191. doi:10.1121/1.1908241 Bourbi, T., Coussy, O., and Zinszner, B., 1987, Acoustics of porous media, Ed. Technip, Paris. Budiansky, B., 1965, On the elastic moduli of some heterogeneous materials: Journal of the Mechanics and Physics of Solids, 13, 223227. doi:10.1016/0022-5096(65)90011-6 Budiansky, B., 1970, Thermal and thermoelastic properties of isotropic composites: Journal Composite Material, 4, 286295. doi:10.1177/ 002199837000400301 Carmichael, R. S., 1989, Practical handbook of physical properties of rocks and minerals: CRC Press, 741 p. Domenico, S. N., 1984, Rock lithology and porosity determination from shear and compressional wave velocity: Geophysics, 49, 11881195. doi:10.1190/1.1441748 Gassmann, F., 1951, ber die elastizitt poroser medien: Vierteljahrsschrift der Naturforschenden Gesellschaft in Zrich, 96, 123. Hill, R. A., 1965, A self-consistent mechanics of composite materials: Journal of the Mechanics and Physics of Solids, 13, 213222. doi:10.1016/00225096(65)90010-4 Japsen, P., Hier, C., Rasmussen, K. L., Fabricius, I., Mavko, G., and Pedersen, J. M., 2002, Effect of uid substitution on ultrasonic velocities on chalk plugs, South Arne eld, North Sea: 72nd Annual International Meeting, SEG, Expanded Abstracts, 21, 18811884. Kerner, E. H., 1956, The elastic and thermoelastic properties of composite media: Proceedings of the Physical Society. Section B, 69, 808813. doi:10.1088/0370-1301/69/8/305 King, M. S., 1966, Wave velocities in rocks as a function of changes in overburden pressure and pore uid saturants: Geophysics, 31, 5073. doi:10.1190/1.1439763 King, M.S., 2005, Rock-physics developments in seismic exploration: a personal 50-year perspective: Geophysics, 70, 3ND8ND. doi:10.1190/1.2107947 Mavko, G., Mukerji, T., and Dvorkin, J., 1998, The rock physics handbook: Tools for seismic analysis in porous media: Cambridge University Press. Palaz, I. and Marfurt, K., J., 1997, Carbonate seismology: Society of exploration geophysicists.

154

Exploration Geophysics

A. Misaghi et al.

Plona, T. J., 1980, Observation of a second bulk compressional wave in a uid-saturated porous solid at ultrasonic frequencies: Applied Physics Letters, 36, 259261. doi:10.1063/1.91445 Prasad, M., and Nur, A., 2003, Velocity and attenuation anisotropy in reservoir rocks: 73rd Annual International Meeting, SEG, Expanded Abstracts, 22, 16521655. Rafavich, F., Kendall, C. H. St. C., and Todd, T. P., 1984, The relationship between acoustic properties and the petrographic character of carbonate rocks: Geophysics, 49, 16221636. doi:10.1190/1.1441570 Sharma, R., Prasad, M., Surve, G., and Katiyar, G. C., 2006, On the applicability of Gassmann model in carbonates: SEG Expanded Abstract,, 25, 18661870. doi:10.1190/1.2369889 Smith, T. M., Sondergeld, C. H., and Rai, C. S., 2003, Gassmann uid substitutions: A tutorial: Geophysics, 68, 430440. doi:10.1190/ 1.1567211

Vo-Thanh, D., 1995, Inuence of uid chemistry on shear-wave attenuation and velocity in sedimentary rocks: Geophysical Journal International, 121, 737749. doi:10.1111/j.1365-246X.1995.tb06435.x Wang, Z., 2001, Fundamentals of seismic rock physics: Geophysics, 66, 398412. doi:10.1190/1.1444931 White, J. E., 1983, Underground sound: Application of seismic waves: Elsevier Science Publishing Co.. Wu, T. T., 1966, The effect of inclusion shape on the elastic moduli of a two-phase material: International Journal of Solids and Structures, 2, 18. doi:10.1016/0020-7683(66)90002-3 Wyllie, M. R. J., Gardner, G. H. F., and Gregory, A. R., 1963, Addendum to studies of elastic wave attenuation in porous media: Geophysics, 28, 1074. doi:10.1190/1.1439306 Manuscript received 25 July 2009; manuscript accepted 17 February 2010.

http://www.publish.csiro.au/journals/eg

You might also like