Estimating the hydrocarbon volume from elastic and resistivity data: A concept

Similar documents
We apply a rock physics analysis to well log data from the North-East Gulf of Mexico

Rock physics of a gas hydrate reservoir. Gas hydrates are solids composed of a hydrogen-bonded ROUND TABLE

ROCK PHYSICS DIAGNOSTICS OF NORTH SEA SANDS: LINK BETWEEN MICROSTRUCTURE AND SEISMIC PROPERTIES ABSTRACT

Downloaded 11/02/16 to Redistribution subject to SEG license or copyright; see Terms of Use at Summary.

Geological Classification of Seismic-Inversion Data in the Doba Basin of Chad*

Effects of fluid changes on seismic reflections: Predicting amplitudes at gas reservoir directly from amplitudes at wet reservoir

Rock physics integration of CSEM and seismic data: a case study based on the Luva gas field.

RESEARCH PROPOSAL. Effects of scales and extracting methods on quantifying quality factor Q. Yi Shen

ROCK PHYSICS MODELING FOR LITHOLOGY PREDICTION USING HERTZ- MINDLIN THEORY

Determination of reservoir properties from the integration of CSEM and seismic data

Rock Physics Perturbational Modeling: Carbonate case study, an intracratonic basin Northwest/Saharan Africa

Pressure and Compaction in the Rock Physics Space. Jack Dvorkin

BPM37 Linking Basin Modeling with Seismic Attributes through Rock Physics

Rock physics and AVO analysis for lithofacies and pore fluid prediction in a North Sea oil field

Rock physics and AVO applications in gas hydrate exploration

The elastic properties such as velocity, density, impedance,

We Density/Porosity Versus Velocity of Overconsolidated Sands Derived from Experimental Compaction SUMMARY

Integration of Rock Physics Models in a Geostatistical Seismic Inversion for Reservoir Rock Properties

Downloaded 02/05/15 to Redistribution subject to SEG license or copyright; see Terms of Use at

Methane hydrate rock physics models for the Blake Outer Ridge

Seismic reservoir and source-rock analysis using inverse rock-physics modeling: A Norwegian Sea demonstration

Towards Interactive QI Workflows Laurie Weston Bellman*

Shaly Sand Rock Physics Analysis and Seismic Inversion Implication

Integrating rock physics and full elastic modeling for reservoir characterization Mosab Nasser and John B. Sinton*, Maersk Oil Houston Inc.

Reservoir Rock Properties COPYRIGHT. Sources and Seals Porosity and Permeability. This section will cover the following learning objectives:

Uncertainties in rock pore compressibility and effects on time lapse seismic modeling An application to Norne field

Calibration of the petro-elastic model (PEM) for 4D seismic studies in multi-mineral rocks Amini, Hamed; Alvarez, Erick Raciel

Downloaded 10/29/15 to Redistribution subject to SEG license or copyright; see Terms of Use at

IDENTIFYING PATCHY SATURATION FROM WELL LOGS Short Note. / K s. + K f., G Dry. = G / ρ, (2)

2011 SEG SEG San Antonio 2011 Annual Meeting 771. Summary. Method

QUANTITATIVE INTERPRETATION

Reducing Uncertainty through Multi-Measurement Integration: from Regional to Reservoir scale

Simultaneous Inversion of Clastic Zubair Reservoir: Case Study from Sabiriyah Field, North Kuwait

Net-to-gross from Seismic P and S Impedances: Estimation and Uncertainty Analysis using Bayesian Statistics

Quantitative Interpretation

GRAIN SORTING, POROSITY, AND ELASTICITY. Jack Dvorkin and Mario A. Gutierrez Geophysics Department, Stanford University ABSTRACT

SEG Houston 2009 International Exposition and Annual Meeting. that the project results can correctly interpreted.

Integrating seismic, CSEM, and well-log data for reservoir characterization

Researcher 2015;7(9)

Sensitivity Analysis of Pre stack Seismic Inversion on Facies Classification using Statistical Rock Physics

ACAE -- ESTUDIO PILOTO DE DIAGNÓSTICO DE FÍSICA DE ROCAS DEL YACIMIENTO CABALLOS EN LOS CAMPOS PUERTO COLON - LORO Y HORMIGA AREA SUR PUTUMAYO

Sections Rock Physics Seminar Alejandra Rojas

Reservoir Characteristics of a Quaternary Channel: Incorporating Rock Physics in Seismic and DC Resistivity Surveys

Keywords. Unconsolidated Sands, Seismic Amplitude, Oil API

AFI (AVO Fluid Inversion)

Integrating rock physics modeling, prestack inversion and Bayesian classification. Brian Russell

Role of Data Analysis in fixing parameters for petrophysics & rockphysics modeling for effective seismic reservoir characterization A case study

LINK BETWEEN ATTENUATION AND VELOCITY DISPERSION

An empirical method for estimation of anisotropic parameters in clastic rocks

Seismic reservoir characterization using the multiattribute rotation scheme: Case study in the South Falkland Basin

Summary. Simple model for kerogen maturity (Carcione, 2000)

Subsurface Consultancy Services

Uncertainty analysis for the integration of seismic and CSEM data Myoung Jae Kwon & Roel Snieder, Center for Wave Phenomena, Colorado School of Mines

Quantitative interpretation using inverse rock-physics modeling on AVO data

Competing Effect of Pore Fluid and Texture -- Case Study

Module for: Resistivity Theory (adapted/modified from lectures in PETE 321 (Jensen/Ayers))

Verification of Archie Constants Using Special Core Analysis and Resistivity Porosity Cross Plot Using Picket Plot Method

P314 Anisotropic Elastic Modelling for Organic Shales

Generation of synthetic shear wave logs for multicomponent seismic interpretation

Seismic reservoir characterization of a U.S. Midcontinent fluvial system using rock physics, poststack seismic attributes, and neural networks

A New AVO Attribute for Hydrocarbon Prediction and Application to the Marmousi II Dataset*

Downloaded 11/20/12 to Redistribution subject to SEG license or copyright; see Terms of Use at

Reservoir And Elastic Property Prediction Away From Well Control Kyle T. Spikes and Jack P. Dvorkin; Stanford Rock Physics Laboratory

From loose grains to stiff rocks The rock-physics "life story" of a clastic sediment, and its significance in QI studies

Th LHR2 08 Towards an Effective Petroelastic Model for Simulator to Seismic Studies

Per Avseth (Dig Science) and Tapan Mukerji (Stanford University)

Crosswell tomography imaging of the permeability structure within a sandstone oil field.

Rock Physics of Shales and Source Rocks. Gary Mavko Professor of Geophysics Director, Stanford Rock Physics Project

The Stanford VI reservoir

Reservoir connectivity uncertainty from stochastic seismic inversion Rémi Moyen* and Philippe M. Doyen (CGGVeritas)

Reservoir Characterization using AVO and Seismic Inversion Techniques

THE ROCK PHYSICS HANDBOOK

Lithology prediction and fluid discrimination in Block A6 offshore Myanmar

Probabilistic seismic inversion using pseudo-wells

water L v i Chapter 4 Saturation

Theoretical Approach in Vp/Vs Prediction from Rock Conductivity in Gas Saturating Shaly Sand

RESERVOIR SEISMIC CHARACTERISATION OF THIN SANDS IN WEST SYBERIA

Reliability of Seismic Data for Hydrocarbon Reservoir Characterization

Statistical Rock Physics

Rock Physics Interpretation of microstructure Chapter Jingqiu Huang M.S. Candidate University of Houston

F003 Geomodel Update Using 4-D Petrophysical Seismic Inversion on the Troll West Field

Introduction. Theory. GEOHORIZONS December 2007/22. Summary

Seismic inversion for reservoir properties combining statistical rock physics and geostatistics: A review

Rock Physics and Quantitative Wavelet Estimation. for Seismic Interpretation: Tertiary North Sea. R.W.Simm 1, S.Xu 2 and R.E.

V Sw =1 * V (1-S w ) 2* (V Sw =1 -V (1-Sw )) * (TWT 99 -TWT 94 ) under reservoir conditions (Rm 3 ) V Sw =1

Dynamic GeoScience Martyn Millwood Hargrave Chief Executive OPTIMISE SUCCESS THROUGH SCIENCE

Pre-Stack Seismic Inversion and Amplitude Versus Angle Modeling Reduces the Risk in Hydrocarbon Prospect Evaluation

An overview of AVO and inversion

We Simultaneous Joint Inversion of Electromagnetic and Seismic Full-waveform Data - A Sensitivity Analysis to Biot Parameter

Downloaded 10/02/18 to Redistribution subject to SEG license or copyright; see Terms of Use at

SEISMIC INVERSION OVERVIEW

Petrophysical Charaterization of the Kwale Field Reservoir Sands (OML 60) from Wire-line Logs, Niger Delta, Nigeria. EKINE, A. S.

CHARACTERIZING RESERVOIR PROPERTIES OF THE HAYNESVILLE SHALE USING THE SELF-CONSISTENT MODEL AND A GRID SEARCH METHOD.

KOZENY-CARMAN EQUATION REVISITED. Jack Dvorkin Abstract

AVO Crossplotting II: Examining Vp/Vs Behavior

GNGTS 2015 Sessione 3.1

The prediction of reservoir

PETROPHYSICAL EVALUATION CORE COPYRIGHT. Petrophysical Evaluation Approach and Shaly Sands Evaluation. By the end of this lesson, you will be able to:

3D petrophysical modeling - 1-3D Petrophysical Modeling Usning Complex Seismic Attributes and Limited Well Log Data

Practical Gassmann fluid substitution in sand/shale sequences

Transcription:

INTERPRETER S CORNER Coordinated by Rebecca B. Latimer Estimating the hydrocarbon volume from elastic and resistivity data: A concept CARMEN T. GOMEZ, JACK DVORKIN, and GARY MAVKO, Stanford University, California Estimating the volume of hydrocarbons in place, the product of porosity and saturation, is a major problem in applied geophysics, but the solution is not straightforward. Reflection seismology depends on the contrasts of the elastic properties (impedance and velocity) that are predominantly affected by porosity and mineralogy. The elastic properties of a rock may also strongly depend on whether it is wet or hydrocarbon-saturated. However, their dependence on hydrocarbon saturation (if larger than zero and smaller than one) is weak. As a result, interpreting seismic data to quantify saturation is often difficult. On the other hand, resistivity strongly reacts to a combination of porosity and saturation, but the latter is impossible to estimate without making an assumption about the former. Consequently, neither seismic nor resistivity data, if analyzed separately, can provide the desired estimate of hydrocarbons in place. The question is: Can these two types of data be used together to address the problem? In this paper, we discuss a physicsdriven solution that combines two theoretical models, one that relates the elastic-wave velocity to porosity, mineralogy, and pore fluid and one that relates resistivity to porosity and saturation. This approach allows us to produce templates of the normalized resistivity versus the P-wave impedance that quantify porosity and saturation. We show examples of such templates for a sandstone reservoir with methane hydrate and with gas. This concept is based on rock physics models and is amenable to the upscaling required to interpret field measurements. We use upscaling to generate a template for the Nuggets gas field. Our model-driven templates are flexible; they can be constructed to honor the site-specific rock properties to account for, for example, diagenetic cementation and depositional sorting as well as the presence of shale. Figure 1. Poisson s ratio (PR) versus the P-wave impedance for unconsolidated clean gas sand, color-coded by the total porosity (left) and water saturation (right), assuming pore pressure=30 MPa, temperature=80 C, and gas gravity=0.65. Figure 2. Water saturation versus the P-wave impedance (left) and versus the normalized resistivity (right), color-coded by the total porosity for the same unconsolidated clean gas sand as in Figure 1. 710 THE LEADING EDGE JUNE 2008

Figure 3. Normalized resistivity versus the P-wave impedance template for the same unconsolidated clean gas sand as in Figure 1. The large red circle is the intersection of the hypothetical resistivity and impedance data with small red circles showing its projections on the porosity and saturation contours. Figure 4. Same as Figure 3 but for cemented sandstone (using the stiff-sand model). Primer on rock physics impedance-resistivity templates. Consider a clean, unconsolidated gas sand whose porosity (φ) varies from 0.1 0.4 and whose water saturation (S w ) is between 0.2 1.0. Figure 1 displays the elastic properties as computed from a soft-sand model (Dvorkin and Nur, 1996), which is appropriate to describe the elastic properties of unconsolidated sand and shale. Note that the impedance (I p ) strongly reacts to φ but only weakly depends on S w, unless S w =1.0. Poisson s ratio (υ) also helps discriminate between wet- and gasfilled sediment but fails to quantify S w. The left side of Figure 2 reaffirms, using the same sand, that S w cannot be quantified from I p ; note the S w contour lines are essentially vertical for S w < 1. The right side of Figure 2 shows the normalized resistivity (R t /R w ), where R t is the measured resistivity and R w is that of water, as calculated from Archie s 1942 resistivity equation. S w strongly reacts to R t /R w ; however, it cannot be quantified without knowing φ. At R t /R w =100, S w can be 0.25 for φ=0.4 or about 0.70 for φ=0.2. Figure 3 shows how the same relationships (between φ, S w, and I p, and between φ, S w, and R=R t /R w ) are used to create a resistivity-impedance mesh. We first fixed porosity (e.g., 0.4), and computed I p and R t /R w using the soft sand model and Archie s equation for water saturations from 1 0.2, generating the first constant porosity line of the mesh on the left. Then, the same procedure was repeated six more times, changing porosity to 0.35, 0.3, 0.25, 0.2, 0.15, and 0.1. Then, we kept water saturation constant (e.g., 1.0) and computed I p and R t /R w using the same models but changing porosity this time from 0.1 0.4, producing the first constant saturation mesh line at the bottom. Again, the rest of the constant saturation mesh lines are computed by repeating this last step eight more times. The intersection of the two measurements can be projected upon the saturation and porosity contours to yield (in this example) φ=0.2 and S w =0.50. Having both measurements helps constrain porosity and saturation; taking each measurement separately yields wide ranges for these variables. The power of this approach is its flexibility; numerous models of rock physics velocity-porosity and resistivity-porosity-saturation are available, meaning site-specific templates can be generated that honor texture (e.g., cemented versus uncemented rock) as well as lithology. For example, Figure 4 displays an I p versus R template in which the soft-sand model is replaced by a stiff-sand model appropriate for contact-cemented rock. As expected, the same impedance and resistivity inputs produce larger porosity and gas saturation than in Figure 3 because at the same porosity, cemented rock is faster than its uncemented counterpart. This illustrates the importance of using the right model for the right rock. This flexibility accommodates not only the rock type but also the fluid, or, more generally, the pore-filling material. For example, in sand containing methane hydrate, both velocity JUNE 2008 THE LEADING EDGE 711

Figure 5. Same as Figure 3 but for a methane hydrate reservoir. The well data (symbols) are color-coded by water saturation (one minus hydrate saturation). Dark blue means high amounts of hydrate. The red and brown symbols are for shale and thus fall outside of the mesh which is built for clean sand. Figure 6. Rock physics diagnostics. Impedance versus porosity for sand and shale in the Nuggets well. The impedance is for 100% watersaturated rock obtained from the original log data by theoretically substituting gas in the reservoir with the formation water. The upper curve is the constant-cement model for 100% quartz while the lower curve is for 100% clay. and resistivity strongly dominate over the background. To build a relevant I p versus R template (Figure 5), we select the elastic and resistivity models describing a hydrate reservoir as established for the Mallik 2L-38 hydrate well (Canada) by Cordon et al. (2006). The log data superimposed upon this rock physics mesh indicate that the porosity of the reservoir is about 0.3 with the hydrate saturation up to 0.8, which is consistent with the values directly observed in the well. Nuggets Field example. The Nuggets gas field is offshore the U. K. in 115 m of water. The sand reservoir belongs to the Eocene Frigg Formation and has a thickness of about 25 m (Harris and MacGregor, 2006). The first step is to use the well-log velocity and porosity data to establish this reservoir as constant-cement sand (a model originally discussed in Avseth et al., 2000), which is consistent with this formation s age (the earliest of the three divi- 712 THE LEADING EDGE JUNE 2008

Figure 7. Impedance-resistivity template for the Nuggets reservoir with log data superimposed. Color code is water saturation (left) and porosity (right). Figure 8. Nuggets gas field well log (blue) and seismic/csem (red) data at the well. The full seismic stack was provided by TGS-NOPEC, from which we obtained the seismic impedance using the Hampson-Russell inversion package. The CSEM resistivity profile was provided by OHM. The reservoir zone is highlighted in green. 714 THE LEADING EDGE JUNE 2008

Figure 9. One realization of a pseudo-well. (left to right) Log-scale impedance; seismic-scale impedance; a synthetic normal-incidence trace; logscale resistivity; and CSEM-scale resistivity versus the two-way traveltime. Red vertical bars show the upscaling windows for the impedance (12.5 m, first frame) and resistivity (150 m, fourth frame). The depth was converted to twt using average velocity in the interval displayed. Figure 10. (left) Log-scale impedance-resistivity template, same as in Figure 7. (right) Seismic/CSEM-scale template. The symbols are from the seismic impedance and CSEM resistivity profiles, shown in Figure 8, at the reservoir. sions of the Tertiary). This diagnosis includes three steps: (1) bring the entire interval to the common-fluid denominator by theoretically making it all wet (fluid substitution from gas to formation water); (2) plot the resulting impedance-porosity data for sand and shale; and (3) superimpose theoretical impedance-porosity curves upon this plot to find which model describes the trends best. Figure 6 shows that the constant cement-sand model is most appropriate for this specific example. The resistivity is modeled using a version of Archie s formula known as the Humble or Winsauer equation (Schön, 1996), which assumes the Archie constants are a=0.62 and m=2.15. The water salinity is 35 000 ppm, which translates to an R w of about 0.1 ohm-m. The resulting I p versus R template (Figure 7, left) gives the sand s porosity at about 0.3 and gas saturation up to 0.9. 716 THE LEADING EDGE JUNE 2008

This template should not be directly used with field seismic and CSEM data whose scales of measurement are vastly different from those in the well and also from each other (Figure 8). These templates have to be altered to honor the statistics of the data, which vary with the scale of measurement. To conduct such upscaling, we need to remember that seismic and CSEM profiling simultaneously provide two values for a single location in the subsurface. What reservoir properties may generate these two responses? One approach to answering this question is forward modeling. We assume that the geometry of the reservoir everywhere is the same as at the well; i.e., we fix the thickness of the gas and wet sand layers as well as the properties of the shale surrounding this sand. Next, we create a pseudo-earth model (or pseudo-well) as shown in Figure 9 where the impedance in the shale is about 5 km/s g/cc, similar to the original well. Then we vary the porosity in the entire sand from 0.15 to 0.40 with a 0.05 step, and, for each porosity step, vary the water saturation in the gas sand from zero to 1.0 with a 0.01 step. This produces a table of porosity and saturation at the scale of the pseudo-well. From this table, we calculate the impedance using the constant cement-sand model and the resistivity for each realization using the Humble equation. One such realization is shown in Figure 9. For upscaling, we use the Backus average with a 12.5- m running window (about a quarter wavelength) to transform the log-scale impedance to the seismic-scale impedance and the arithmetic average with a 150-m running window to transform the log-scale resistivity to the CSEM-scale resistivity. The latter window (150 m) was selected to honor the difference between the resistivity at the reservoir as recorded by logging and CSEM. The upscaled impedance and resistivity profiles are shown in Figure 9. We select those upscaled values in the middle of the reservoir for each realization and create an R t /R w versus I p mesh which yields a template usable at the field scale (Figure 10). To interpret data from Nuggets Field for porosity and saturation, we take the seismically derived impedance and CSEM resistivity selected from the depth interval around the reservoir. These values are placed upon the field-scale template in Figure 10. The contour lines in this template indicate that the gas saturation is around 0.85 and porosity is around 0.25, consistent with the actual values in Nuggets. Conclusions. By combining two theoretical models, one that relates the elastic properties of rock to porosity, mineralogy, and pore fluid, and the other that relates resistivity to porosity and saturation, we create templates of the normalized resistivity versus P-wave impedance. We can generate these templates at the well log-scale and also upscale them to the seismic/csem scales of measurement. Such upscaled templates can be directly used with field data to obtain simultaneous estimates of porosity and saturation. The answer is not unique. Many parameters can vary away from well control, including the geometry as well as the shale and reservoir properties. For an exhaustive interpretation, they all have to be varied within reasonable ranges and probability distributions. This will eventually yield not a single answer but probabilistic distributions of porosity and saturation. We envision that the approach discussed in this paper may produce useful invariants, such as porosity times saturation times thickness (or net-to-gross), which should be relied upon in reserve estimates. This is a subject of future work where the existing arsenal of stochastic modeling can be used within the framework presented here. Finally, we have to reiterate that the rock physics models for the impedance-resistivity templates should be selected to reflect the geologic nature of rock both on the elastic and resistivity sides (the latter by, for example, using the Waxman-Smits-Juhasz equation instead of the Archie equation). Also, instead of considering just the P-wave impedance, other seismic attributes can be explored, such as the elastic impedance and attenuation, to further constrain the estimates. Suggested reading. The electrical resistivity log as an aid in determining some reservoir characteristics by Archie (Transactions of AIME, 1942). Rock physics diagnostics of North Sea sands: Link between microstructure and seismic properties by Avseth et al. (Geophysical Research Letters, 2000). Seismic reflections of gas hydrate from perturbational forward modeling by Cordon et al. (GEOPHYSICS, 2006). Elasticity of high-porosity sandstones: Theory for two North Sea data sets by Dvorkin and Nur (GEOPHYSICS, 1996). On the universality of diagenetic trends by Dvorkin et al. (TLE, 2002). Lithology substitution in fluvial sand by Dvorkin et al. (TLE, 2004). Determination of reservoir properties from the integration of CSEM and seismic data by Harris and MacGregor (First Break, 2006). Gas Hydrate Estimation Error Associated with Uncertainties of Measurements and Parameters by Lee and Collett (U. S. Geological Survey Bulletin, 2001). Physical Properties of Rocks: Fundamentals and Principles of Petrophysics: Handbook of Geophysical Exploration by Schön (Pergamon Press, 1996). TLE Acknowledgments: Thanks to Lucy MacGregor and Peter Harris of OHM/RSI for their advice. Thanks to OHM/RSI and TGS-NOPEC for providing the data. Thanks to Hampson-Russell for providing their software to Stanford University. This work was supported by the Stanford Rock Physics and Borehole Geophysics Project, and the DOE Basic Energy Sciences Grant DE-FG02-03ER15423. Corresponding author: gomezct@stanford.edu 718 THE LEADING EDGE JUNE 2008