Effects of Surfactant-Emulsified Oil-Based Mud on Borehole Resistivity Measurements

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

SATURATION-HEIGHT AND INVASION CONSISTENT HYDRAULIC ROCK TYPING USING MULTI-WELL CONVENTIONAL LOGS

IMPROVED ESTIMATION OF MINERAL AND FLUID VOLUMETRIC CONCENTRATIONS IN THINLY-BEDDED AND INVADED FORMATIONS

SPE These in turn can be used to estimate mechanical properties.

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

COPYRIGHT PETROSKILLS LLC

WETTABILITY CHANGE TO GAS-WETNESS IN POROUS MEDIA

THE UNIVERSITY OF TRINIDAD & TOBAGO

Q f is the flow rate of mud filtrate across the

IMPROVED ASSESSMENT OF IN-SITU FLUID SATURATION WITH MULTI-DIMENSIONAL NMR MEASUREMENTS AND CONVENTIONAL WELL LOGS

FORMATION EVALUATION PETE 321

Evaluation of Rock Properties from Logs Affected by Deep Invasion A Case Study

6. THE BOREHOLE ENVIRONMENT. 6.1 Introduction. 6.2 Overburden Pressures

PETROPHYSICAL EVALUATION CORE COPYRIGHT. Saturation Models in Shaly Sands. By the end of this lesson, you will be able to:

Petrophysical Rock Typing: Enhanced Permeability Prediction and Reservoir Descriptions*

INFERRING RELATIVE PERMEABILITY FROM RESISTIVITY WELL LOGGING

Well Logging. Salam Al Rbeawi 2011

MONTE CARLO MODELING OF NUCLEAR MEASUREMENTS IN VERTICAL AND HORIZONTAL WELLS IN THE PRESENCE OF MUD- FILTRATE INVASION AND SALT MIXING

Basics of Geophysical Well Logs_Porosity

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

SPE Copyright 2013, Society of Petroleum Engineers

N121: Modern Petrophysical Well Log Interpretation

Well Logging Importance in Oil and Gas Exploration and Production

The sensitivity of the array resistivity log to mud. inversion for improved oil water recognition

SPE Pore-Level Analysis of the Relationship Between Porosity, Irreducible Water Saturation, and Permeability of Clastic Rocks

Well Log Simulation- A new vista in Formation Evaluation

EXTENDED ABSTRACT EVALUATING THE SHALY SAND OIL RESERVOIRS OF EL TORDILLO FIELD, ARGENTINA, USING MAGNETIC RESONANCE LOGS

SPE Copyright 2012, Society of Petroleum Engineers

INTRODUCTION TO LOGGING TOOLS

Copyright. Ankur Gandhi

Improved estimation of mineral and fluid volumetric concentrations from well logs in thinly bedded and invaded formations

Evaluation of Petrophysical Properties of an Oil Field and their effects on production after gas injection

The role of capillary pressure curves in reservoir simulation studies.

Inversion-based detection of bed boundaries for petrophysical evaluation with well logs: Applications to carbonate and organic-shale formations

Comparison of Classical Archie s Equation with Indonesian Equation and Use of Crossplots in Formation Evaluation: - A case study

Calculation of Irreducible Water Saturation (S wirr ) from NMR Logs in Tight Gas Sands

CORING TO CONSERVE RESIDUAL WATER SATURATION AND ITS USE WHEN INTERPRETING ELECTRICAL LOGS

GEOLOGICAL LOG INTERPRETATION TUTORIAL

SPE =. (3) h=0.7035p c / ρ (1)

MOVEMENT OF CONNATE WATER DURING WATER INJECTION IN FRACTURED CHALK

Permeability Restoration of Saudi Arabia rocks. Osama A. El-Mahdy

GEOHORIZONS. Ankur Gandhi, Carlos Torres-Verdín, Ben Voss, Federico Seminario Gros, and Johnny Gabulle

Introduction to Formation Evaluation Abiodun Matthew Amao

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

water L v i Chapter 4 Saturation

Technology of Production from Shale

NEW SATURATION FUNCTION FOR TIGHT CARBONATES USING ROCK ELECTRICAL PROPERTIES AT RESERVOIR CONDITIONS

Fundamentals Of Petroleum Engineering FORMATION EVALUATION

CO 2 Foam EOR Field Pilots

Study on the Four- property Relationship of Reservoirs in YK Area of Ganguyi Oilfield

Evaluation of Low Resistivity Laminated Shaly Sand Reservoirs

Petrophysics. Theory and Practice of Measuring. Properties. Reservoir Rock and Fluid Transport. Fourth Edition. Djebbar Tiab. Donaldson. Erie C.

Lalaji Yadav* and Troyee Das Reliance Industries Limited, Navi Mumbai, , India

Geological and Petrophysical Evaluation for an Oil Well

Field Scale Modeling of Local Capillary Trapping during CO 2 Injection into the Saline Aquifer. Bo Ren, Larry Lake, Steven Bryant

Fracture relative permeability revisited

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

Saturation Modelling: Using The Waxman- Smits Model/Equation In Saturation Determination In Dispersed Shaly Sands

FORMATION EVALUATION OF SIRP FIELD USING WIRELINE LOGS IN WESTERN DEPOBELT OF NIGER DELTA

Permeability Estimates & Saturation Height Functions: A talk of two halves. Dr Joanne Tudge LPS Petrophysics 101 Seminar 17 th March 2016

Exploration / Appraisal of Shales. Petrophysics Technical Manager Unconventional Resources

COPYRIGHT PETROSKILLS LLC

FIELD-STUDY OF INTEGRATED FORMATION EVALUATION IN THINLY LAMINATED RESERVOIRS

Plumbing the Depths of the Pelican Field

RELATIONSHIP BETWEEN CAPILLARY PRESSURE AND RESISTIVITY INDEX

Oil and Natural Gas Corporation Limited, 4th Floor GEOPIC, Dehradun , Uttarakhand

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

ECS Elemental Capture Spectroscopy Sonde. Fast, accurate lithology evaluation

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

Determination of the Laminar, Structural and Disperse Shale Volumes Using a Joint Inversion of Conventional Logs*

Log Interpretation Parameters Determined from Chemistry, Mineralogy and Nuclear Forward Modeling

An empirical method for estimation of anisotropic parameters in clastic rocks

SPE * Mark of Landmark (Halliburton Energy Services). Copyright 2002, Society of Petroleum Engineers Inc.

Neutron Log. Introduction

True 3D measurements for enhanced reservoir quantification. Rt Scanner

PORE-SCALE OBSERVATIONS OF THE EFFECT OF SURFACE TREATED NANOPARTICLES ON DRAINAGE PROCESSES

DETERMINING THE SATURATION EXPONENT BASED ON NMR PORE STRUCTURE INFORMATION

1: Research Institute of Petroleum Industry, RIPI, Iran, 2: STATOIL ASA, Norway,

New Generation Permeability Logs

Geologic influence on variations in oil and gas production from the Cardium Formation, Ferrier Oilfield, west-central Alberta, Canada

Petrophysical Analysis and Sequence Stratigraphy Appraisal from Well logs of Bobo field, South-Eastern, Niger Delta

Formation Evaluation: Logs and cores

Risk Factors in Reservoir Simulation

Field Examples of the Combined Petrophysical Inversion of Gamma-ray, Density, and Resistivity Logs acquired in Thinly-bedded Clastic Rock Formations 1

LITTLE ABOUT BASIC PETROPHYSICS

SRC software. Rock physics modelling tools for analyzing and predicting geophysical reservoir properties

Understanding NMR. Geoff Page Baker Hughes Region Petrophysics Advisor Baker Hughes Incorporated. All Rights Reserved.

Geologic and Reservoir Characterization and Modeling

GY 402: Sedimentary Petrology

DNAPL migration through interbedded clay-sand sequences

SPE Copyright 2006, Society of Petroleum Engineers

Mixed Reservoir Wetting in Unconventional Reservoirs and Interpretation of Porosity/Resistivity Cross Plots, Derived From Triple-combo Log Data

Correlation Between Resistivity Index, Capillary Pressure and Relative Permeability

COMBINED SIMULATION AND INVERSION OF SP AND RESISTIVITY LOGS FOR THE ESTIMATION OF CONNATE WATER RESISTIVITY AND ARCHIE S CEMENTATION EXPONENT

Site Characterization & Hydrogeophysics

Nuclear Magnetic Resonance Log

Effect of Dielectric Permittivity Distribution in Invaded Zone on Induction Log Data

SCA : A STRUCTURAL MODEL TO PREDICT TRANSPORT PROPERTIES OF GRANULAR POROUS MEDIA Guy Chauveteau, IFP, Yuchun Kuang IFP and Marc Fleury, IFP

Appraising a late-middle-aged Brent Group field

Abstract. Introduction to Area

Transcription:

Effects of Surfactant-Emulsified Oil-Based Mud on Borehole Resistivity Measurements Jesús M. Salazar,* SPE, Carlos Torres-Verdín, SPE, and Gong Li Wang,** SPE, The University of Texas at Austin Summary We quantify the influence of oil-based mud (OBM)-filtrate invasion and formation-fluid properties on the spatial distribution of fluid saturation and electrical resistivity in the near-wellbore region. The objective is to appraise the sensitivity of borehole resistivity measurements to the spatial distribution of fluid saturation resulting from the compositional mixing of OBM and in-situ hydrocarbons. First, we consider a simple two-component formulation for the oil phase (OBM and reservoir oil) wherein the components are first-contact miscible. A second approach consists of adding water and surfactant to a multicomponent OBM invading a formation saturated with multiple hydrocarbon components. Simulations also include presence of irreducible, capillary-bound, and movable water. The dynamic process of OBM invasion causes component concentrations to vary with space and time. In addition, the relative mobility of the oil phase varies during the process of invasion because oil viscosity and oil density are both dependent on component concentrations. Presence of surfactants in the OBM is simulated with a commercial adaptive implicit compositional formulation that models the flow of three-phase multicomponent fluids in porous media. Simulations of the process of OBM invasion yield 2D spatial distributions of water and oil saturation that are transformed into spatial distributions of electrical resistivity. Subsequently, we simulate the corresponding array-induction measurements assuming axial-symmetric variations of electrical resistivity. We perform sensitivity analyses on field measurements acquired in a well that penetrates a clastic formation and that includes different values of density and viscosity for mud filtrate and formation hydrocarbon. These analyses provide evidence of the presence of a high-resistivity region near the borehole wall followed by a low-resistivity annulus close to the noninvaded resistivity region. Such an abnormal resistivity annulus is predominantly caused by high viscosity contrasts between mud filtrate and formation oil. The combined simulation of invasion and array-induction logs in the presence of OBM invasion provides a more reliable estimate of water saturation, which improves the assessment of in-place hydrocarbon reserves. Introduction In general, OBM increases drilling rates and provides better-quality boreholes than when drilling with water-based mud (WBM). In OBMs, the continuous fluid phase is a mixture of liquid hydrocarbons. This continuous phase dominates the process of invasion and mixes with formation fluids. Water is also present in the form of an emulsion. Chemical emulsifiers/surfactants are added in the mud to prevent water droplets from coalescing and leaving the emulsion (Bourgoyne et al. 1986) and to ensure that the weighting material is wetted by oil (Skalli et al. 2006). Calcium or magnesium fattyacid soaps are typically used as emulsifiers. *Now with ConocoPhillips. **Now with Schlumberger. Copyright 2011 Society of Petroleum Engineers This paper (SPE 109946) was accepted for presentation at the SPE Annual Technical Conference and Exhibition, Anaheim, California, 11 14 November 2007, and revised for publication. Original manuscript received for review 27 July 2007. Revised manuscript received for review 31 August 2010. Paper peer approved 8 September 2010. Array-induction resistivity measurements are invariably acquired in OBM environments. Such measurements exhibit several radial lengths of investigation, typically from 10 to 90 in. and, in some cases, 120 in. The variability of resistivity curves exhibiting different lengths of investigation is, in most cases, a footprint of mud-filtrate invasion. When array-induction measurements are acquired in the presence of WBM, large separation of resistivity curves and deep invasion can be expected because of a monotonic change of formation resistivity in the radial direction. Ideally, we would not expect such a behavior in the presence of OBM. However, in hydrocarbon zones, OBM filtrate can replace native hydrocarbon and movable water, thereby resulting in an invasion front with different resistivity values in the near-wellbore region (La Vigne et al. 1997). Moreover, invading mud filtrate is miscible with native oil. In the mixing process, OBM causes changes of fluid density and fluid viscosity, thereby modifying the apparent oil-phase mobility in the invaded zone (Malik et al. 2008). The objective of this paper is to quantify the effect of fluid properties (density and viscosity) and composition on arrayinduction measurements acquired in the presence of OBM-filtrate invasion. Special attention is given to the presence of surfactants in the OBM. Previous laboratory experiments with core samples (Van et al. 1988; Yan and Sharma 1989) show that surfactants greatly change the wettability of sandstone cores from a strongly water-wet to an oil-wet condition. Most of these effects were inferred from changes of contact angle that caused a reduction of interfacial tension. Our study is centered on measurements acquired in a North Sea well penetrating clastic rock formations. The well was drilled with OBM, whereas resistivity measurements were acquired with the array-induction imager tool (AIT) (Schlumberger). We select three depth intervals that exhibit distinct invasion behavior at different depth intervals in the formation. Additionally, we give a detailed background of the method used to simulate the process of OBM-filtrate invasion and the associated array-induction resistivity measurements. A base case with high initial water saturation is used to perform several sensitivity analyses that quantify the influence of formation-fluid properties and mud components. A second example allows us to perform sensitivity analyses with a light-hydrocarbon depth interval subject to invasion with OBM that includes an emulsified surfactant. The third example includes a gas-saturated interval close to irreducible water saturation. Finally, we perform the simulation of OBM-filtrate invasion in a long depth interval encompassing a complete capillary transition zone from light oil at the top to water at the bottom. Geological Description and Depositional System The field case considered in this work corresponds to a Paleocene sandstone dome located in the central North Sea (Martin et al. 2005). The field is composed of an average of 60 m of gas cap with a gas/oil ratio of 871 scf/bbl and 58 m of oil-saturated column with a gravity of 40 API followed by an active aquifer (BP 2003). Preserved Paleocene strata in the North Sea show siliciclastic sediments with minor amounts of coal, tuff, volcaniclastic rocks, marls, and reworked carbonate sediments (Ahmadi et al. 2003). During the early Paleocene, submarine fans developed in the faulted basin margins in the west, spread out, and overlapped to form near-continuous sandstone bodies. The specific formation under analysis is predominantly composed of noncalcareous, blocky gray mudstones interbedded with sandy, high-density gravity-flow 608 September 2011 SPE Journal

10000 1000 Perm r=0.1 r=0.5 r=2 r=10 [µm] k/φ [md/p.u.] 5000 1000 Permeability, md 100 10 1 0.1 200 100 50 25 10 50 2.5 1.0 0.5 0.25 0.05 0.01 0 5 10 15 20 25 30 Porosity, % Fig. 1 Crossplot of core permeability vs. core porosity along with parametric curves of Windland s critical pore-throat radius, r 35, (continuous colored lines) and k/ ratio (gray and dashed-black lines). Most of the samples are located in zones of high porosity and high permeability, which indicates the high flow and storage capacity of the rock units. deposits and minor volcaniclastic rocks. An estimated 35% of net sand volume makes this formation one of the more sand-prone systems of the Paleocene formations in the region. Rock units within the formation consist of alternations of fine-grained to medium-grained or, in some cases, coarse-grained sandstone with common mudstone and chalk clasts. Porosity ranges between 20 and 28%, while permeability k varies from 50 md in low-porosity zones to 1000 md in high-porosity intervals. Fig. 1 is a plot of core permeability vs. core porosity, showing Windland s r 35 (critical pore-throat radius) curves (Pittman 1992) along with k/ isolines. This plot indicates that the formation consists of high-quality rock units with pore-throat radii greater than 2 m and k/ ratios greater than 200 md/p.u. Petrophysical Analysis of Field Measurements The upper section of the formation corresponds to gas-saturated interbedded sandstones. In the intermediate section, the well penetrates thick sand bodies partially oil-saturated toward a capillary transition zone. The lower depth interval of the formation reaches an active aquifer. Routine core measurements of porosity and permeability are available for the complete hydrocarbon interval and for several locations within the water zone. This well provides a TABLE 1 SUMMARY OF ARCHIE S PARAMETERS AND ROCK AND FLUID PROPERTIES USED TO ESTIMATE WATER SATURATION AND POROSITY Variable U nits Value Archie s tortuosity factor a 1.00 Archie s cementation exponent m 1.89 Archie s saturation exponent n 1.92 Connate water resistivity at 240 F m 0.12 Matrix density Shale density Water density g/cm g/cm g/cm 3 3 3 2.655 2.55 1.00 Hydrocarbon density (gas zone mix) g/cm 3 0.60 Hydrocarbon density (oil zone mix) g/cm 3 0.75 textbook example of an ideal hydrocarbon reservoir. It shows a gas cap at the top, followed by an oil column penetrating a capillary transition zone toward the free-water/oil contact (WOC). Water Saturation. Relatively clean sandstones with low water saturation S w at the top and high S w downward compose the formation. Therefore, we used Archie s equation (Archie 1942) to calculate S w : m 1 n z Sw r, z R ( r, z) = ( ) ( ),.......................... (1) a R t w where r is radial distance from the wellbore; z is vertical distance with respect to the top of the formation; R t is true formation resistivity; R w is connate-water resistivity; m and n are Archie s cementation and saturation exponents, respectively; and a is the tortuosity factor. When a rock becomes oil-wet in the invaded zone, n is no longer constant in the radial direction (Donaldson and Siddiqui 1989; Dawe and Grattoni 1998). However, resistivity-index measurements with wettability alteration were not available for this study, whereby we assumed n to be constant across the interval of analysis. Table 1 describes the input parameters used to calculate initial water saturation using Archie s equation. Such parameters were obtained from laboratory measurements performed on core samples and fluid samples withdrawn from the formation. Shaliness, Porosity, and Permeability. In order to diagnose the type of shale distribution across the formation, we applied Thomas and Stieber s method (Thomas and Stieber 1975) to gamma ray measurements and to computed total porosity. Fig. 2 shows the Thomas-Stieber crossplot indicating that dispersed shale is the dominant type of shale distribution in the formation under analysis. Routine core measurements of porosity and permeability are available in the well. Therefore, we calculated core-calibrated curves of porosity and permeability. Nonshale porosity was calculated from density and neutron measurements by accounting for the presence of two fluids (oil/gas and water) and two minerals (quartz and clay) in the porous medium. Permeability was calculated by means of a modified Timur-Tixier equation from porosity and irreducible water saturation (Salazar et al. 2006). We specialized the equations for specific intervals of the formation with calibration of porosity/permeability measurements. The coefficient and September 2011 SPE Journal 609

Total Porosity, fraction 0.3 0.25 0.2 0.15 Laminated Dispersed Hydrocarbon Components. Original formation hydrocarbons consist of components in the range from C 1 through C 18+. For Case 1, we assume only liquid hydrocarbons and lump them into a single component (namely, C 7 C 18 ). In Case 2, we add a lighter component, C 4 C 6, which, in turn, is the most abundant in the zone. For the last case (Case 3), we include a gaseous component (N 2 C 1 ), which is the leading component in that depth interval. Regarding OBM, we assume that it consists of hydrocarbons in the range of C 14 through C 18, which are lumped into a single MC 16 component. The binary-interaction parameter between the hydrocarbon pseudocomponents is assumed null. Under dynamic drilling conditions, drilling mud mixes with solid particles and formation fluids as the drill bit penetrates the formation. This behavior can further modify the composition of the mud. 0.1 90 80 70 60 50 40 30 20 10 Gamma Rays, γapi Fig. 2 Thomas-Stieber crossplot used to diagnose the type of shale/clay distribution in the formation. Dispersed shale is the dominant type of shale/clay distribution in the sand units. exponents of the equations were estimated by means of multilinear least-squares regression. Fig. 3 shows the results obtained from petrophysical analysis. We observe that the water-saturation curve resembles a typical capillary pressure curve as a function of depth. Cases of Study. Three different intervals with distinct invasion behavior were chosen from the cored well to perform simulations of the process of OBM-filtrate invasion and to simulate the corresponding resistivity measurements. Each case was considered a single-layer homogeneous formation. Case 1 is a 16-ft interval within the oil zone ( 8,616 ft) and was used to calibrate our synthetic base case. In such an interval, variability of array-induction resistivity curves indicates presence of an invasion profile. This variability allows us to perform sensitivity analyses to several fluid and formation properties. By manually matching the simulated apparent resistivity curves to field measurements, we assess the parameters for capillary pressure and relative permeability to be used in the remaining depth intervals along the well. Case 2 is composed of a depth interval between gas and oil. We observe a 7-ft interval that exhibits invasion (8,519.5 8,526.5 ft). The latter is considered a very-light-hydrocarbon zone. Case 2 is used to perform additional sensitivity analyses in light oil with different values of initial water saturation and presence of emulsifier surfactants in the OBM. Additionally, Case 3 corresponds to a gas zone at the top of the formation (8,329 8,345 ft) with initial water saturation close to irreducible water saturation. Therefore, when OBM-filtrate invades the formation, all three phases (invading mud, formation water, and gas) are mobile in the near-wellbore region. Mud filtrate is assumed an emulsion of hydrocarbons, water, and surfactants. Table 2 summarizes the average petrophysical properties used in the simulations. Fig. 4 shows well logs (gamma ray, caliper, resistivity, density, and neutron) for each case where we can diagnose the corresponding invasion behavior. Finally, we consider a very long depth interval (8,527 8,806 ft) that includes the end of the gas zone, the oil zone, and the aquifer. The objective is to model the complete capillary transition zone using a simple binary OBM-filtrate invasion model and to study the corresponding effect on array-induction resistivity measurements. Composition of Fluids With the main objective of performing sensitivity analyses of fluid properties, we assume different compositions of both formation oil and mud filtrate across different depth intervals in the formation. We also consider presence of irreducible and movable water in these depth zones. Surfactants. Presence of water emulsified with surfactants is also considered in the sensitivity analyses. Common surfactants in the industry are known by their commercial names EZ-MUL, X-VIS, and INVERMUL (Baroid/Halliburton), among others. The viscosity of these surfactants may vary from less than 1 cp to slightly more than 1000 cp. Most of such surfactants act as emulsifiers and oil-wetting agents. They are composed of organic, fatty-acid soaps. When the soap forms a mixture with water and oil, soap molecules accumulate at oil/water interfaces, with the water-soluble extreme residing in the water phase and the oil-soluble extreme residing in the oil phase (Bourgoyne et al. 1986). In this paper, we assume a generic surfactant (henceforth referred to as EMUL) composed by stearic acid (La Scala et al. 2004; Lide 2006), which is a chain of 18 carbon atoms. Table 3 summarizes the pseudoproperties of both the lumped hydrocarbon components and the emulsifying surfactant (Lide 2006). Numerical Simulation of the Process of OBM-Filtrate Invasion We approach the simulation of the process of OBM-filtrate invasion assuming axial-symmetric formation properties with respect to the axis of a vertical borehole. Simulations enforce boundary and source flow-rate conditions on specific depth segments along the wellbore. The outer, lower, and upper limits of the formation consist of impermeable zones with no-flow boundary conditions (Aziz and Settari 2002). Simulation cases consider both waterfree and water-emulsified OBM. Therefore, salt concentration is assumed constant across the formation and equal to that of connate water. A commercial adaptive-implicit formulation simulator known as STARS (Computer Modelling Group) is used to perform the simulations and the sensitivity analysis of fluid properties. The STARS software is an advanced three-phase multicomponent fluid-flow simulator with the capability of simulating the flow of emulsions and foams in porous media (STARS 2006). With this simulator, we are able to treat the OBM filtrate as a multicomponent fluid containing oil and water emulsified within the oil with surfactants. Formation hydrocarbon is also considered a multicomponent fluid in which we account for the presence of gas. The geometry used to simulate all the cases of study consists of a finite-difference grid with nodes logarithmically spaced in the radial direction (50 nodes). The wellbore radius is equal to 0.51 ft, and the external radius is 2,000 ft. A logarithmic radial grid allows us to properly reproduce the rapid space/time variations of pressure and component concentrations in the near-borehole region. Along the vertical direction, grid nodes are uniformly spaced at 0.5-ft intervals. We used the parametric equations proposed by Brooks and Corey (1966) to define capillary pressure P c and relative permeability k ri curves (Delshad and Pope 1989). Maximum capillary pressure was assigned from interfacial tension between reservoir fluids by means of Washburn s equation (Showalter 1979). Laboratory core data were not available for oil/water or gas/liquid systems. Therefore, typical exponents and coefficients for this type of formation were used in the simulation. In selecting the Brooks-Corey parameters, we also accounted for the fact that rock wettability was altered (Yan 610 September 2011 SPE Journal

Fig. 3 Petrophysical assessment. Track 1: gamma ray and caliper logs. Track 2: array-induction resistivity logs. Track 3: neutrondensity logs highlighting light-hydrocarbon zones (shown with shading) and photoelectric-factor log. Track 4: calculated effective porosity (shown with a continuous line) and core porosity (shown with circles). Track 5: modified Timur-Tixier permeability (shown with a continuous line) and core permeability (shown with circles). Track 6: Archie s water saturation. Track 7: volumetric analysis indicating clean sands in the oil and water zones and more-heterogeneous sands in the gas zone. TABLE 2 SUMMARY OF ASSUMED PETROPHYSICAL PROPERTIES FOR THE CASES OF STUDY Property U nits Case 1 Case 2 Case 3 Thickness Effective porosity Water saturation ft fraction fraction 16 0.25 0.363 6 0.27 0.230 Volumetric shale concentration fraction 0.04 0.03 0.07 Horizontal permeability md 353 434 446 Vertical permeability md 180 217 223 16 0.2 0.18 September 2011 SPE Journal 611

8,330 8,335 8,340 Caliper, in. Neutron, fraction 12 14 16 18 20 22 Case No. 3 0.4 0.2 0 8,345 0 50 100 150 30 50 100 130 2 2.2 2.4 2.6 2.8 Gamma Ray, γapi Bulk Density, g/cm 3 Caliper, in. Neutron, fraction 12 14 16 18 20 22 Case No. 2 0.4 0.2 0 8,518 8,519 8,520 8,521 8,522 8,523 8,524 8,525 8,526 8,527 0 50 100 150 10 15 20 30 40 50 2 2.2 2.4 2.6 2.8 Gamma Ray, γapi Bulk Density, g/cm 3 Caliper, in. Neutron, fraction 12 14 16 18 20 22 Case No. 1 0.4 0.2 0 8,602 8,604 8,605 8,608 8,612 8,614 GR CAL AT10 AT20 AT30 AT60 AT90 RHOB NPHI 8,616 0 50 100 150 10 15 20 25 2 2.2 2.4 2.6 2.8 Gamma Ray, γapi Bulk Density, g/cm 3 Fig. 4 Well logs for the three depth intervals used to perform sensitivity analysis of OBM-filtrate invasion to fluid properties. The center panels display 2-ft-resolution AIT apparent-resistivity curves. Logs were acquired in a partially oil-saturated interval (Case 1), a very-light-hydrocarbon zone (Case 2), and a partially gas-saturated sand (Case 3). Measurements include three fluid-saturation levels of a capillary transition zone. and Sharma 1989; Skalli et al. 2006) when invading the formation with OBM. Initially, the rock formation under analysis was assumed water-wet. As the OBM invasion started, rock wettability changed to preferentially oil-wet. Hence, the critical value of saturation where relative permeability of oil is equal to that of water was less than 50%. Table 4 summarizes the parameters used in the Brooks-Corey model for the simulations of OBM-filtrate invasion considered in this paper. Appendix A describes in detail the capillary pressure and relative permeability equations used in the simulations. Flow Rate of Invasion. In the process of OBM-filtrate invasion, the flow rate q mf is a rather uncertain parameter. It depends both on mud properties and on formation properties (permeability, porosity, water saturation, capillary pressure, and relative permeability). We simulated array-induction resistivity measurements in a field base case, assuming that the formation was invaded with WBM. Once we reproduced the resistivity measurements, we used the average flow rate of WBM invasion as the initial estimate of flow rate of invasion for the simulation of OBM-filtrate invasion (Wu et al. 2005; Malik et al. 2008; Salazar and Torres-Verdín 2009). Subsequently, we adjusted the rate as needed to reproduce borehole resistivity measurements. This procedure yielded physically consistent flow rates of OBM-filtrate invasion. Table 5 describes the additional formation and fluid properties necessary to simulate the process of OBM-filtrate invasion. Simulation of Induction Resistivity Measurements. From the simulation of OBM-filtrate invasion, we obtained spatial distributions (radial and vertical directions) of water saturation. Subsequently, we used Archie s equation to calculate electrical resistivity at each gridblock, assuming that electrical resistivity of formation water was not altered during the process of invasion. Finally, we simulated array-induction resistivity measurements from the spatial distribution of electrical resistivity. In order to perform the simulation of resistivity measurements, we assumed 2D axial symmetry of electrical resistivity, where current loop sources are located at the center of the borehole. We used the numerical mode matching method (NMM) to simulate electrical-resistivity measurements (Chew et al. 1984; Zhang et al. 1999). In summary, the algorithm was initialized with the fluid-flow simulation of OBM invading porous media, which was coupled to the NMM code to simulate borehole resistivity measurements. Field Studies We began the simulation of mud-filtrate invasion with Case 1. The objective was to estimate the initial flow rate of invasion by matching array-induction resistivity measurements. Table 2 shows the average petrophysical properties used as inputs to the simulation. Calibration Case. At the outset, we assumed that the invasion took place with fresh WBM and calculated the flow rate of invasion (Wu et al. 2005). Then, we used the calculated initial flow rate as input to the simulation of OBM-filtrate invasion (Malik et al. 2008). To that end, we assumed that the formation was saturated with movable and irreducible water and liquid hydrocarbons lumped into the single component C 7 C 18. On the other hand, OBM-filtrate was described with hydrocarbons in the range of C 14 through C 18 lumped into a generic MC 16 component. This fluid mixture is often referred to as a binary component mixture. Table 3 describes the properties of both formation and filtrate fluids. In order to secure a good match between field and simulated resistivity measurements, we modified capillary pressure, relative permeability, and flow rate of invasion and performed the simulation of invasion several times. While modifying the latter parameters, we accounted for changes of wettability undergone by the rock formation while being invaded with OBM. Table 4 describes the parameters for capillary pressure and relative permeability used to match resistivity measurements. Fig. 5 describes the spatial distributions (radial and vertical directions) of water saturation and electrical resistivity calculated after 60 hours of OBM-filtrate invasion. Fig. 6 compares field to simulated AIT resistivity curves calculated after the simulation of invasion along with the average horizontal permeability and matching flow rate of invasion. Simulated array-induction resistivity measurements agree well with field measurements. The average flow rate necessary to reproduce the measured apparent resistivity curves was approximately 0.24 (ft 3 /D)/ft with a radial length of invasion approximately equal to 1.8 ft. This flow rate was used as the initial flow rate (q 0 mf ) of invasion throughout the sensitivity 612 September 2011 SPE Journal

TABLE 3 SUMMARY OF PRESSURE/VOLUME/TEMPERATURE PROPERTIES OF THE ASSUMED IN-SITU HYDROCARBON AND MUD-FILTRATE COMPONENTS Property Units N 2 C 1 C 4 C 6 C 7 C 18 EMUL MC 16 Critical temperature F 125.7 359.8 656.2 986.7 822.5 Critical pressure psi 653.3 498.2 322.2 188.6 240.2 Mass density lbm/ft 3 1.05 12.06 40.2 59.9 45.7 Molar weight lbm/mol 16.6 67.73 132.8 284.5 222 Viscosity cp 0.025 0.105 1.0 20 1.5 Variable TABLE 4 SUMMARY OF RELATIVE PERMEABILITY AND CAPILLARY PRESSURE PARAMETERS USED IN BROOKS-COREY EQUATIONS V alue Empirical exponent for wetting phase, e w 3 Empirical exponent for nonwetting phase, e nw 7 End point for wetting phase, 0.80 End point for nonwetting phase, 1 Empirical exponent for pore-size distribution, e p 4 Capillary pressure coefficient,, [psi D 1/2 ] 8, 1.2* Interfacial Tension, cos [dynes/cm] 26, 4* * When OBM included surfactants. analyses. We remark that the previous simulations assumed a single-layer homogeneous formation. For the sensitivity analyses, we considered the presence of surfactants in the OBM. Moreover, gaseous hydrocarbons were included as pseudocomponents of the formation fluid. Fig. 7 shows the radial distribution of simulated density, viscosity, water saturation, and electrical resistivity obtained with the STARS simulator. Sensitivity Analysis. We used STARS to perform analyses of the sensitivity of the process of invasion to fluid properties for each 8,602 8,604 8,606 8,608 8,612 8,614 Water Saturation, fraction (a) 0.38 0.37 0.36 0.35 0.34 0.33 0.32 Formation 0.31 1 2 3 4 1 2 3 4 Fig. 5 Spatial distributions (vertical and radial directions) of water saturation (a) and electrical resistivity (b) calculated after 60 hours of OBM-filtrate invasion into a partially oil-saturated formation (Case 1). (b) 20 19 18 17 16 15 14 Variable TABLE 5 SUMMARY OF PARAMETERS ASSUMED IN THE SIMULATION OF THE PROCESS OF MUD-FILTRATE INVASION U nits Value Initial formation pressure at 8,616 ft psi 3,628 Maximum invasion time Initial invasion flow rate Temperature at 8,616 ft Formation outer boundary hours case based on different assumptions. First, we quantified the sensitivity of apparent resistivity measurements to native-oil viscosity and density. Second, we made perturbations of native-oil viscosity and capillary pressure while assuming OBM to be an emulsion that included oil, water, and surfactants. The first two sensitivity analyses were performed for Case 1 only. Finally, we moved upward in the formation (Cases 2 and 3) and performed sensitivity analyses to initial water saturation and flow rate of invasion in multicomponent light-oil and gas zones subject to invasion with emulsified OBM filtrate. An additional case of study consisted of simulating a long capillary transition zone using a simple binary formulation for compositional fluid mixing. All the simulations assumed 60 hours of invasion. Sensitivity to Formation-Oil Viscosity and Mass Density. We simulated Case 1 using a simple binary formulation for compositional fluid mixing. For the base case, viscosity of native-oil pseudocomponent C 7 C 18 was equal to 1.0 cp and mass density was equal to 40.2 lbm/ft 3. The OBM-filtrate MC 16 viscosity was equal to 1.5 cp, and density was equal to 45.7 lbm/ft 3. Each perturbation for a given property was performed separately. Formation-oil viscosity was changed to 15 and to 0.3 cp, whereas density was modified to 55.5 and to 25.5 lbm/ft 3 in that order. Fig. 8 shows radial distributions of electrical resistivity calculated for each perturbation. We observe that resistivity increases close to the borehole wall, while becoming asymptotic toward R t (initial watersaturation conditions) radially deeper into the formation. This behavior is because of the relatively large volume of movable water that is displaced by mud filtrate. Irreducible water saturation was approximately 8%, and initial water saturation was approximately 36%. Therefore, the large capillary transition zone causes a radially smooth invasion front. Fig. 9 shows the corresponding simulated AIT measurements for each perturbation. The most remarkable conclusion of this analysis is that, for the case of higher viscosity, mud-filtrate moves faster into the formation, thereby increasing the electrical resistivity in the near-wellbore region. Thus, we observe larger separation of apparent-resistivity curves than for the case of no perturbation. Changes of fluid density do not entail a measurable difference on the radial length of invasion. Likewise, array-induction resistivity curves are less affected by gravity in the presence of high-density hydrocarbons than in the presence of lowdensity hydrocarbons, as indicated by their symmetric behavior within a particular layer. 60 ft 3 /d/ft 0.2375 F ft 213 2,000 Residual water saturation fraction 0.08 Residual oil saturation fraction 0.10 September 2011 SPE Journal 613

2-ft Resolution Permeability, md 0 200 400 8,602 q mf 8,604 8,605 8,608 R10-field R20-field R30-field R60-field R90-field R10-sim R20-sim R30-sim R60-sim R90-sim k h 8,612 8,614 8,616 10 15 20 25 (a) 0 0.2 0.4 Flow Rate, (ft 3 /D)/ft (b) Fig. 6 (a) Comparison of simulated (continuous thick curves) and measured (circles and thin curves) 2-ft array-induction resistivity measurements after 60 hours of OBM-filtrate invasion for Case 1. (b) Assumed average flow rate per foot and horizontal permeability. The assumed constant ratio k v /k h 0.5 causes gravity-segregation effects, which give rise to a vertically asymmetric resistivity profile. Oil Density, lbm/ft 3 46 45 44 43 42 41 OBM-Filtrate Formation oil Oil Viscosity, cp 1.6 1.5 1.4 1.3 1.2 1.1 1 OBM-Filtrate Formation oil 40 1 2 3 4 5 0.9 1 2 3 4 5 0.4 20 Water Saturation, fraction 0.38 036 0.34 0.32 0.3 1 2 3 4 5 18 16 14 12 1 2 3 4 5 Fig. 7 Radial distributions (clockwise from top left) of oil density, oil viscosity, electrical resistivity, and water saturation for the base case (Case 1) after 60 hours of invasion. 614 September 2011 SPE Journal

32 30 28 26 24 22 20 18 16 14 Base Case μ o =15.0 cp μ o =0.3 cp ρ o =55.5 lbm/ft 3 ρ o =25.5 lbm/ft 3 1 1.5 2 2.5 3 3.5 4 4.5 5 Fig. 8 Radial distributions of electrical resistivity obtained after 60 hours of invasion for Case 1 by performing separate perturbations of formation-oil viscosity and density. The base case (blue circles) assumed native-oil viscosity of 1 cp and density of 40.2 lbm/ft 3. Variations in the viscosity contrast between native oil and filtrate produced significant changes in the near-wellbore resistivity (red stars and black squares for 15 and 0.3 cp, respectively). However, the effect of density contrast was negligible on resistivity measurements. Sensitivity to Oil/Water Interfacial Tension and Formation- Oil Viscosity With Emulsified OBM. Case 1 was simulated assuming that OBM contained water and emulsifying surfactant EMUL. We assumed that water and EMUL were two new components of the original OBM. The molar fraction for each component was 0.1 water, 0.4 EMUL, and 0.5 MC 16. Viscosities were 1 cp for water and 20 cp for the surfactant. This analysis was performed to quantify the decrease of interfacial tension times contact angle ( cos ) because of the presence of surfactants in the OBM (Skalli et al. 2006) at two different values of formation-oil viscosity (15 and 0.3 cp). The base case assumed a water-wet system with cos = 26 dynes/cm (maximum capillary pressure equal to 6.7 psi). On the other hand, for the highly oil-wet case, we assumed cos = 4 dynes/cm (maximum capillary pressure equal to 1.0 psi). Fig. 10 shows radial distributions of water saturation and electrical resistivity associated with each perturbation. Fig. 11 displays the AIT resistivity measurements simulated from the spatial distributions of electrical resistivity. For the case of high interfacial tension, we observe the same behavior as in the preceding example. However, electrical resistivity is slightly higher in this case. Lower interfacial tension causes a remarkable increase of electrical resistivity near the borehole wall. Such a behavior is because of the presence of surfactants in the OBM, which decrease the influence of capillary forces because of the decrease of interfacial tension. Sensitivity to Initial Water Saturation and Flow Rate of Invasion With Emulsified OBM. Case 2 was the base case for this analysis. We used the same OBM composition as in the preceding example. The formation was saturated with very- light-hydrocarbon pseudocomponents (namely C 4 C 6 and C 7 C 18 with molar 8,605 8,615 Base Case R10 R20 R30 R60 R90 Base Case C 7 C 18 μ o =1.0 cp ρ o =40.2 lbm/ft 3 μ o =15.0 cp ρ o =55.5 lbm/ft 3 8,605 8,615 8,605 μ o =0.3 cp ρ o =25.5 lbm/ft 3 8,615 13 15 17 20 25 30 13 15 17 20 25 30 Fig. 9 Sensitivity analysis of array-induction resistivity measurements to changes of formation-oil viscosity and density. Simulated 2-ft AIT resistivity measurements for Case 1 are compared to the corresponding measurements simulated from the sensitivity analysis. A large viscosity contrast caused large radial variability of resistivity measurements. The oil-density contrast did not cause radial variability but gave rise to a vertically asymmetric resistivity profile for low oil density because of gravity segregation. September 2011 SPE Journal 615

Water Saturation, fraction 0.4 0.35 0.3 0.25 0.2 70 60 Base Case σcosθ=26 dynes/cm, μ o =15 cp σcosθ=26 dynes/cm, μ o =0.3 cp σcosθ=4 dynes/cm, μ o =15 cp σcosθ=4 dynes/cm, μ o =0.3 cp 50 40 30 20 10 1 1.5 2 2.5 3 3.5 4 4.5 5 Fig. 10 Radial distributions of water saturation and electrical resistivity obtained after 60 hours of invasion for Case 1. Mud filtrate consists of MC 16, EMUL, and water. The radial distributions describe simulation results obtained by performing joint perturbations of interfacial tension and native-oil viscosity. Low interfacial tension and large viscosity contrast caused the largest variability (magenta triangles) observed on the radial resistivity distribution. Base Case Surfactant Base Case No Surfactant 8,605 8,615 σcosθ=26 dynes/cm, μ o =15 cp σcosθ=4 dynes/cm, μ o =15 cp 8,605 8,615 R10 R20 R30 R60 R90 σcosθ=26 dynes/cm, μ o =0.3 cp σcosθ=4 dynes/cm, μ o =0.3 cp 8,605 8,615 10 20 30 40 50 10 20 30 40 50 Fig. 11 Sensitivity analysis of array-induction resistivity measurements to changes of both oil/water interfacial tension and formation-oil viscosity. The Base Case 1 (with and without presence of surfactants) simulated AIT resistivity measurements are compared to simulations obtained from the sensitivity analysis with a surfactant emulsifying the water present in the OBM. 616 September 2011 SPE Journal

8,520 8,522 8524 Water Saturation, fraction 1 2 3 4 5 1 2 3 4 5 (a) 0.28 0.26 0.24 0.22 Formation fractions of 0.7 and 0.3, respectively). Initial water saturation S win was equal to approximately 23% and capillary pressure was low, similar to the low-interfacial-tension case of the preceding example. We changed S win to 18 and 28% and flow rate to 0.96 (4q 0 mf ) and 0.06 (q 0 mf /4) (ft 3 /D)/ft. This case was also simulated 0.2 0.18 Fig. 12 Spatial distributions (vertical and radial directions) of water saturation (a) and electrical resistivity (b) calculated after 60 hours of OBM-filtrate invasion into a partially light-oilsaturated formation (Case 2). A low-resistivity annulus arises approximately 1.5 ft away from the borehole. (b) 50 40 30 20 assuming a water- and surfactant-free OBM. Fig. 12 describes the spatial distributions of water saturation and electrical resistivity calculated after 60 hours of OBM-filtrate invasion for Base Case 2. Water saturation increased above S win at approximately 1.5 ft away radially from the borehole, which was responsible for the presence of a low-resistivity annulus in the zone. This annulus is more evident in Fig. 13, where we show radial distributions of water saturation and electrical resistivity separately for each perturbation. With a surfactant-free mud, the annulus vanishes and is similar to the case of very low flow rate of invasion with emulsified OBM. Fig. 14 shows the corresponding AIT resistivity measurements simulated from the spatial distributions of electrical resistivity. Larger separation of apparent-resistivity curves is observed in the cases of high flow rate of invasion and high initial water saturation than in the cases of low S win and low q mf. In general, presence of surfactants causes larger variability of the simulated apparentresistivity curves. Sensitivity to Flow Rate of Invasion With Emulsified OBM Invading a Gas-Bearing Formation. In Case 3, fluids saturating the formation consist mainly of gas and very light oil. Mole fractions of gaseous pseudocomponents were 0.7 N 2 C 1 and 0.3 C 4 C 6. On the other hand, the mole fraction of liquid components was 0.5 for both C 4 C 6 and C 7 C 18, respectively. Case 3 was located in the upper depth section of the formation and its initial water saturation (18%) was close to irreducible water saturation (8%). Capillary pressure parameters were the same as those assumed in the preceding example. We changed the flow rate of invasion to 0.96 (4q 0 mf ) and 2.4 (10q 0 mf ) (ft 3 /D)/ft. Fig. 15 shows the spatial distributions of water saturation and electrical resistivity calculated after 60 hours 0 of 10 q mf of mud-filtrate invasion. Presence of a low-resistivity annulus was noteworthy as well as deep invasion because of the exaggerated flow rate. Fig. 16 shows radial distributions of water saturation and electrical resistivity associated with each perturbation. The higher the flow rate is, the lower the resistivity in the low-resistivity-annulus region is and the deeper its radial location Water Saturation, fraction 0.35 0.3 0.25 0.2 70 60 50 40 30 Base Case No Surf. S win =28% S win =18% q mf =4q o mf q mf =4q o mf /4 20 10 1 1.5 2 2.5 3 3.5 4 4.5 5 Fig. 13 Radial distributions of water saturation and electrical resistivity obtained after 60 hours of invasion for Case 2. Mud filtrate for the base case consists of MC 16, EMUL, and water. The plots show simulation results obtained by performing separate perturbations of initial water saturation and flow rate of invasion. A low-resistivity annulus arises when the OBM includes surfactants. September 2011 SPE Journal 617

8,518 Base Case Surfactant Base Case No Surfactant 8,520 8,522 8,524 8,526 8,518 8,520 8,522 8,524 8,526 S win =28% R10 R20 R30 R60 R90 o q mf =4q mf 8,518 S win =18% o q mf =4q mf /4 8,520 8,522 8,524 8,526 10 20 30 40 50 10 20 30 40 50 Fig. 14 Sensitivity analysis of array-induction resistivity measurements to changes of initial water saturation and flow rate of invasion. Simulated AIT resistivity measurements for Base Case 2 (with and without presence of surfactants) are compared to those obtained from the sensitivity analysis with a surfactant emulsifying the water in the OBM. is away from the wellbore. For the simulation with water- and surfactant-free OBM, the annulus zone was negligible. Fig. 17 shows the simulated AIT apparent-resistivity measurements where we observe large separation of resistivity curves caused by the 8,330 8,332 8,334 8,336 8,338 8,340 8,342 8,344 Water Saturation, fraction (a) Formation 0.205 0.2 0.195 0.19 0.185 0.18 0.175 0.17 0.165 0.16 1 2 3 4 5 1 2 3 4 5 Fig. 15 Spatial distributions (vertical and radial directions) of water saturation and electrical resistivity calculated after 60 hours of OBM-filtrate invasion into a partially gas-saturated formation (Case 3). The flow rate of invasion was exaggerated to emphasize the low-resistivity annulus. (b) 100 90 80 70 higher flow rate of invasion. The difference between simulated AIT measurements with and without the presence of surfactants was not significant. Such a behavior was because of the low initial water saturation under a low flow rate of invasion; the resistivity annulus was relatively shallow, thus not affecting the array-induction resistivity measurements. Simulation of Invasion in a Capillary Transition Zone. The simulation in the capillary transition zone (8,527 8,806 ft) assumed the same conditions as those of the calibration case that is to say, single OBM-filtrate and native-oil pseudocomponents (binary formulation) and the same parameters for capillary pressure and relative permeability. This example was constructed using 12 petrophysical layers with vertical grids spaced at approximately 1.9-ft intervals. Most of the layers were very thick (> 15 ft) with four thin layers (< 5 ft) to account for sands and intercalated shales. Fig. 18 shows the spatial distributions of water saturation and electrical resistivity calculated after 60 hours of mud-filtrate invasion. These distributions indicated that water saturation increased toward the bottom of the formation within the capillary transition zone. Fig. 19 shows the corresponding measured and simulated AIT resistivity curves along with curves of layer-by-layer flushed-zone and virgin-zone resistivity obtained from the simulation of OBM-filtrate invasion. We observed separation of apparent-resistivity curves in the thicker sands, with the separation more pronounced in the water region at the bottom of the formation. Discussion We studied the influence of several fluid properties on arrayinduction resistivity measurements in a well drilled with OBM. In the base case, we observed that water saturation decreased in the near-wellbore region because of the presence of OBM filtrate invading the formation. The reduction of S w entailed an increase of electrical resistivity. Sensitivity analyses to native-oil viscosity and density indicated that viscosity exerted a measurable control on the 618 September 2011 SPE Journal

Water Saturation, fraction 0.21 0.2 0.19 0.18 0.17 0.16 0.15 120 Base Case Base Case No Surf. o q mf =10q mf q mf =4q o mf 110 100 90 80 70 1 1.5 2 2.5 3 3.5 4 4.5 5 Fig. 16 Radial distributions of water saturation and electrical resistivity obtained after 60 hours of invasion for Case 3. Mud filtrate for the base case consists of MC 16, EMUL, and water. The plots show simulation results obtained by performing separate perturbations of the flow rate of invasion. A low-resistivity annulus arises when the OBM contains surfactants. spatial distribution of water saturation, thereby affecting the spatial distribution of electrical resistivity. The effect of fluid density on water saturation and resistivity was relatively less significant on the simulated apparent-resistivity curves. In the case of large viscosity of native oil, invasion caused a high-electrical-resistivity zone in the near-wellbore region because of the reduction of fluid mobility of the native oleic phase. Therefore, because mud-filtrate mobility was higher than that of formation fluids, OBM filtrate penetrated faster into the porous medium. When emulsifying surfactants were added to the OBM, we observed a similar behavior. However, when we decreased the oil/water interfacial tension, electrical resistivity abruptly increased in the radial zone close to the wellbore. In low-interfacial-tension environments, capillary forces opposing the invasion decreased and, in some cases, caused reduction of residual saturation and increased relative permeability (Dawe and Grattoni 1998), thereby allowing the OBM filtrate to move more freely into 8,330 Base Case Surfactant Base Case No Surfactant 8,335 8,340 8,345 8,330 o q mf =10q mf R10 R20 R30 R60 R90 q mf =4qo mf 8,335 8,340 8,345 60 70 80 90 100 110 60 70 80 90 100 110 Fig. 17 Sensitivity analysis of array-induction resistivity measurements to changes of flow rate of invasion. Simulated AIT resistivity measurements are compared to those obtained from the sensitivity analysis with a surfactant emulsifying the water in the OBM. September 2011 SPE Journal 619

Water Saturation, fraction Formation 8,550 0.9 20 0.8 10 8,650 0.7 0.6 6 8,700 0.5 2 8,750 0.4 1 8,800 0.3 0.6 1 2 3 4 1 2 3 4 Fig. 18 Spatial distributions (vertical and radial directions) of water saturation (a) and electrical resistivity (b) calculated after 60 hours of OBM-filtrate invasion into a formation partially oil saturated toward the top and fully water saturated at the base. The spatial distributions of water saturation and electrical resistivity were constructed assuming a subsurface model that emphasizes vertical changes of resistivity within a capillary fluid transition zone toward the WOC. the formation (La Vigne et al. 1997). In general, alteration of the rock s wettability may occur, depending on time of invasion and the volume and chemical composition of surfactants added to the OBM. In many cases, wettability does not change for very short times of invasion or if the volumetric fraction of surfactants in the mud is too low. However, experience with enhanced-oil-recovery processes that make use of surfactants indicates that wettability alteration may occur after 2 or 3 days of injection. 8,550 8,650 8,700 R xo R10 R20 R30 R60 R90 R t 8,750 8,800 (a) 10 0 10 1 10 0 10 1 (b) Fig. 19 Measured 2-ft-vertical-resolution array-induction resistivity curves compared to their simulated values after resistivity matching performed by manually changing both flow rate of invasion and relative permeability. The panels show layer-by-layer values of flushed-zone (R xo ) and virgin-zone resistivity (R t ) obtained from the simulation of OBM-filtrate invasion. 620 September 2011 SPE Journal

Water Saturation, fraction 1 0.8 0.6 0.4 0.2 0 80 Gas Oil Water 60 40 20 0 1 1.5 2 2.5 3 3.5 4 4.5 5 Fig. 20 Radial distributions of water saturation and electrical resistivity obtained after 60 hours of invasion for three distinct fluids saturating a formation with the same petrophysical properties but different values of initial water saturation. Mud filtrate for the base case consists of MC 16, EMUL, and water. A low-resistivity annulus arises when the OBM includes surfactants and invades a formation with a low-viscosity saturating fluid. The example of a very-light-hydrocarbon zone allowed us to perform additional sensitivity analyses to assess the effect of the flow rate of invasion and saturation zones on the simulated apparent-resistivity curves. We assumed low oil/water interfacial tension and presence of emulsified OBM (which, in general, exhibits a larger viscosity than a single-component drilling fluid). When OBM filtrate invaded the formation, it rapidly displaced the movable water, thereby creating a radial zone with S w greater than S win. The large viscosity contrast between the light hydrocarbon and the OBM together with the low initial water saturation created the necessary conditions to form a radial water bank. In turn, the latter water bank was responsible for the presence of an electrically conductive annulus in the invaded zone. This annulus affected the behavior of the AIT resistivity measurements. The corresponding effect on resistivity measurements depended on the magnitude of the flow rate of invasion and on initial water saturation. Large values of q mf and S win entailed a significant separation of arrayinduction resistivity curves. We noted that oil saturation abruptly increased in the near-wellbore region, which became an oil-wet rock, whereupon Archie s saturation exponent n was no longer constant in the radial direction as reported by other researchers (Donaldson and Siddiqui 1989; Dawe and Grattoni 1998). However, the lack of experimental measurements prompted us to use a constant n across the formation. In our analysis, we observed that, when surfactants were not present in the OBM, a resistivity annulus did not arise. On the other hand, when the formation was gas bearing, the annulus was more pronounced, even within low-water-saturation zones. The radial location of the annulus depended on the flow rate of invasion: The larger the flow rate was, the deeper the radial location of the annulus was, thereby causing larger variability of array-induction resistivity curves. In some cases, the intermediatesensing resistivity curves were sensitive to the annulus region, as indicated by their relatively low values compared to those of the deepest-sensing curves. On the basis of the field study and the sensitivity analyses described earlier, we constructed a final synthetic example of a 16-ft clastic formation with nonshale porosity equal to 25%, irreducible water saturation equal to 8%, horizontal permeability equal to 446 md, and vertical permeability equal to 223 md. The objective was to perform simulations of invasion of emulsified OBM in the formation at different water-saturation conditions. Invasion flow rate was assumed equal to 0.96 (ft 3 /D)/ft (4q 0 mf ). Initially, the top of the formation was assumed in the gas zone (8,329 ft, S win = 18%), then in an oil zone ( ft, S win = 36%), and finally in the aquifer (8,770 ft, S win = 100%); formation temperature and pressure were modified accordingly. Fig. 20 shows radial distributions of water saturation and electrical resistivity for each value of S win, where we observe that the higher the initial water saturation was, the more discontinuous the radial invasion front was. We also observed that the viscosity contrast between emulsified OBM and native hydrocarbon smoothes the radial invasion front. Low native-fluid viscosity and low S win in gas zones are responsible for an electrically conductive annulus, as indicated by the radial distribution of electrical resistivity. Fig. 21 shows the corresponding simulations of AIT resistivity measurements. We observed the largest relative separation of curves between the deepest-sensing (R90) and the shallowest-sensing (R10) AIT resistivity measurements in the water zone, whereas the smallest curve separation was observed within the partially gas-saturated zone. On the other hand, the partially oil-saturated depth interval with movable water exhibited the most-pronounced separation of intermediate-sensing AIT resistivity curves (R20, R30, and R60). The sensitivity analysis showed that, even though AIT resistivity curves in the gas zone did not exhibit large separation among themselves, the presence of a resistivity annulus was more important for AIT measurements acquired in depth zones farther from irreducible water saturation. For completeness, the same simulation example was performed assuming no invasion. We found no separation of array-induction resistivity September 2011 SPE Journal 621

8,329 8,345 8,616 R10 R20 R30 R60 R90 S wir S win 8,770 8,786 0 0.2 0.4 0.6 0.8 1 10 0 10 1 10 2 (a) Water Saturation, fraction Fig. 21 Simulated array-induction resistivity measurements across three distinct fluid zones in a clastic formation. The assumed OBM filtrate includes an emulsifying surfactant. (a) Brown blocks identify shale layers, whereas dark blue, light blue, green, and red identify irreducible-water, movable-water, oil, and gas zones, respectively. (b) curves, thereby confirming that the separation of resistivity curves responded only to invasion and not to processing artifacts associated with the calculation of apparent resistivities. Conclusions Simulated spatial distributions of water saturation were highly influenced by the assumed properties of OBM and native-hydrocarbon components. Such a situation caused abnormally high resistivity near the borehole wall, which was governed by: Capillary forces (in turn modified because of wetting properties of surfactants) Native- and invading-fluid viscosity, thereby affecting relative fluid mobility Saturation zone (initial water saturation and residual fluid saturation) Flow rate of invasion We showed that fluid properties of drilling mud and native hydrocarbon affected apparent-resistivity measurements during overbalanced drilling conditions. The process of OBM-filtrate invasion with emulsifying surfactants caused changes in the spatial distribution of electrical resistivity 10 to 15% larger than when invading with a water- and surfactant-free OBM, thereby causing measurable variability of the simulated AIT resistivity curves. Furthermore, the radial length of invasion decreased on the order of 45 to 55%. A radial water bank ensued when hydrocarbon viscosity was below 0.25 cp together with an absolute difference between S win and S wir less than 25%, giving rise to an electrically conductive annulus in the invaded zone. The latter annulus was more pronounced in gas-bearing than in oil-bearing zones. Initial water saturation became an exceedingly important parameter to reproduce apparent-resistivity measurements in the presence of OBM-filtrate invasion, which allowed us to obtain a more reliable estimate of R t, thereby improving the assessment of in-place hydrocarbon saturation. Moreover, as in the case of WBM-filtrate invasion, presence of an annulus (George et al. 2004) and large separation of resistivity curves indicated that the invaded formation was porous and permeable, possibly with good production potential. The formation under analysis was a clastic sequence with high-porosity and high-permeability rock units. We warn that conclusions stemming from this paper may need to be adjusted for low-porosity and low-permeability formations. Low-porosity rock formations often entail large capillary pressure effects and radially deeper invasion, thereby resulting in significant variability of apparent-resistivity curves. Acknowledgments We are obliged to the Andrew Field Partner Group for providing the field data and for permission to publish the results of our work. We are also thankful to Bob Brugman with Computer Modelling Group for his help in the compositional simulation with STARS. The work reported in this paper was funded by The University of Texas at Austin s Research Consortium on Formation Evaluation, jointly sponsored by Anadarko, Aramco, Baker Atlas, BHP Billiton, BP, British Gas, ConocoPhillips, Chevron, Eni E&P, ExxonMobil, Halliburton Energy Services, Hydro, Marathon Oil Corporation, Mexican Institute for Petroleum, Occidental Petroleum Corporation, Petrobras, Schlumberger, Shell International E&P, Statoil, Total, and Weatherford. References Ahmadi, M., Sawyers, M., Kenyon-Roberts, S., Stanworth, B., Kugler, K., Kristensen, J., and Fugelli, E. 2003. Paleocene. In The Millenium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. D. Evans, C. Graham, A. Armour, and P. Bathurst, Chap. 14, 235 259. London: The Geological Society of London. Archie, G.E. 1942. The Electrical Resistivity Log as an Aid in Determining Some Reservoir Characteristics. SPE-942054-G. Trans., AIME, 146: 54 62. Aziz, K. and Settari, A. 2002. Petroleum Reservoir Simulation. Calgary, Alberta: Blitzprint, Ltd. Bourgoyne, A.T., Chenevert, M.E., and Millheim, K.K. 1986. Principles of BHA. In Applied Drilling Engineering, Vol. 2, 426 443. Richardson, Texas: Textbook Series, SPE. 622 September 2011 SPE Journal

BP. 2003. Asset Portfolio: Andrew Field, http://www.bp.com/liveassets/ bp_internet/globalbp/staging/global_assets/downloads/u/uk_asset_ andrew.pdf (accessed 12 July 2007). Brooks, R.H. and Corey, A.T. 1966. Properties of Porous Media Affecting Fluid Flow. J. Irrigation and Drainage Engineering 92 (IR2): 61 68. Chew, W.C., Barone, S., Anderson, B., and Henessy, C. 1984. Diffraction of Axisymmetric Waves in a Borehole by Bed Boundary Discontinuities. Geophysics 49 (10): 1586 1595. doi: 10.1190/1.1441567. Dawe, R.A. and Grattoni, C.A. 1998. The Visualization of the Pore-Scale Physics of Hydrocarbon Recovery from Reservoirs. First Break 16 (11): 371 386. doi: 10.1046/j.1365-2397.1998.00697.x. Delshad, M. and Pope, G.A. 1989. Comparison of the Three-Phase Oil Relative Permeability Models. Transport in Porous Media 4 (1): 59 83. doi: 10.1007/BF00134742. Donaldson, E.C., and Siddiqui, T.K. 1989. Relationship Between the Archie Saturation Exponent and Wettability. SPE Form Eval 4 (3): 359 362. SPE-16790-PA. doi: 10.2118/16790-PA. George, B.K., Torres-Verdín, C., Delshad, M., Sigal, R., Zouioueche, F., and Anderson, B., 2004. Assessment of In-Situ Hydrocarbon Saturation in the Presence of Deep Invasion and Highly Saline Connate Water. Petrophysics 45 (2): 141 156. La Scala, J.J., Sands, J.M., Orlicki, J.A., Palmese, G.R., and Robinette, E.J. 2004. Fatty Acid-Based Monomers as Styrene Replacements for Liquid Molding Resins. Polymer 45 (22): 7729 7737. La Vigne, J., Barber, T.D., and Bratton, T. 1997. Strange Invasion Profiles: What Multiarray Induction Logs Can Tell Us About How Oil-Based Mud Affects The Invasion Process and Wellbore Stability. Paper 1997-B presented at the SPWLA 38th Annual Logging Symposium, Houston, 15 18 June. Lide, D.R. ed. 2006. CRC Handbook of Chemistry and Physics, 87th edition. Boca Raton, Florida: CRC Press. Malik, M., Salazar, J.M., Torres-Verdín, C., Wang, G.L., Lee, H.J., and Sepehrnoori, K. 2008. Influence of Petrophysical Properties on Array- Induction Measurements Acquired in the Presence of Oil-Base Mud- Filtrate Invasion. Petrophysics 49 (1): 74 92. Martin, A.J., Robertson, D., Wreford, J., and Lindsay, A. 2005. High- Accuracy Oriented Perforating Extends the Sand-Free Production Life of Andrew Field. Paper SPE 93639 presented at Offshore Europe, Aberdeen, 6 9 September. doi: 10.2118/93639-MS. Pittman, E.D. 1992. Relationship of Porosity and Permeability to Various Parameters Derived from Mercury Injection Capillary Pressure Curve for Sandstones. AAPG Bulletin 76 (2): 191 198. Salazar, J.M. and Torres-Verdín, C. 2009. Quantitative Comparison of Processes of Oil- and Water-Based Mud-Filtrate Invasion and Corresponding Effects on Borehole Resistivity Measurements. Geophysics 74 (1): E57 E73. doi: 10.1190/1.3033214. Salazar, J.M., Torres-Verdín, C., Alpak, F.O., Habashy, T.M., and Klein, J.D. 2006. Estimation of Permeability from Borehole Array Induction Measurements: Application to the Petrophysical Appraisal of Tight-Gas Sands. Petrophysics 47 (6): 527 544. Showalter, T.T. 1979. Mechanics of Secondary Hydrocarbon Migration and Entrapment. AAPG Bulletin 63 (5): 723 760. Skalli, L., Buckley, J.S., Zhang, Y., and Morrow, N.R. 2006. Surface and Core Wetting Effects of Surfactants in Oil-Based Drilling Fluids. J. Pet. Sci. Eng. 52 (1 4): 253 360. doi: 10.1016/j.petrol.2006.03.012. STARS User Guide. 2006. Calgary, Alberta: Computer Modelling Group (CMG). Thomas, E.C. and Stieber, S.J. 1975. The Distribution of Shale in Sandstones and its Effects Upon Porosity. Presented at the SPWLA 16th Annual Logging Symposium, New Orleans, 4 7 June. Van, J.-N., Menezes, J.L., and Sharma, M.M. 1988. Wettability Alteration Due to Interactions with Oil-Based Muds and Mud Components. Paper SPE 18162 presented at the SPE Annual Technical Conference and Exhibition, Houston, 2 5 October. Wu, J., Torres-Verdín, C., Sepehrnoori, K., and Proett, M.A. 2005. The Influence of Water-Base Mud Properties and Petrophysical Parameters on Mudcake Growth, Filtrate Invasion, and Formation Pressure. Petrophysics 46 (1): 14 32. Yan, J. and Sharma M.M. 1989. Wettability Alteration and Restoration for Cores Contaminated with Oil-Based Muds. J. Pet. Sci. Eng. 2 (1): 63 76. doi: 10.1016/0920-4105(89)90051-X. Zhang, G.J., Wang, G.L., and Wang, H.M. 1999. Application of Novel Basis Function in a Hybrid Method Simulation of the Response of Induction Logging in Axisymmetrical Stratified Media. Radio Science 34 (1): 19 26. Appendix A In a capillary-tube experiment, capillary pressure P c between two immiscible fluids is given by the Washburn equation (Showalter 1979): P = 2 cos c C,.................................(A-1) r where is the contact angle between the fluid interface and the solid surface in degrees, is interfacial tension between wetting and nonwetting fluids (e.g., water and oil) in dynes/cm, r is the radius of the capillary tube in µm, and C is a unit conversion constant equal to 0.145 psi m/(dynes/cm). For the base case in this paper, we assume a water/oil system with = 30 dynes/cm, = 30, and r = 1.1 m (at irreducible water saturation), which gives a maximum capillary pressure equal to 6.7 psi at reservoir conditions. The radius of the capillary tube is equivalent to the pore-throat aperture. We further use the Brooks-Corey relationship (Brooks and Corey 1966; Delshad and Pope 1989) to calculate saturation-dependent capillary pressure and relative permeabilities. The relationship for capillary pressure is given by 0 e Pc = Pc ( 1 S p n ),............................(A-2) k where P 0 c is the coefficient for capillary pressure, e p is the pore-size distribution exponent, is porosity, k is permeability expressed in darcies, and S n is normalized wetting-phase saturation, given by Sw Swr Sn =,..............................(A-3) 1 Swr Snwr where S w and S wr are saturation and irreducible saturation, respectively, for the wetting phase, and S nwr is irreducible saturation for the nonwetting phase. According to the Brooks-Corey model, wetting-phase relative permeability k rw is given by k e = k 0 S w,...................................(a-4) rw rw n where k 0 rw is wetting-phase endpoint relative permeability and e w is an empirical exponent for the wetting phase. The nonwetting-phase relative permeability k rnw is given by ( ) 0 k = k 1 S,.............................(A-5) rnw rnw n e nw 0 where k rnw is nonwetting-phase endpoint relative permeability and e nw is an empirical exponent for the nonwetting phase. Jesús M. Salazar is a petrophysicist in the Subsurface Technology group of ConocoPhillips. Since January 2011, he has worked in the Australia business unit in Perth, Australia. Before joining ConocoPhillips, Salazar worked for Petróleos de Venezuela and for the Center for Petroleum Engineering at The University of Texas at Austin. He holds PhD and MS degrees in petroleum engineering from The University of Texas at Austin and a BS degree in physics with honors from Universidad Central de Venezuela. Salazar is interested in formation evaluation, inverse problems, and near-wellbore fluid-flow simulation. He is a technical reviewer for the SPE Journal, the SPE Reservoir Evaluation & Engineering journal, and Geophysics and received an award for the best paper published in Petrophysics in 2006. Carlos Torres-Verdín holds a PhD degree in engineering geoscience from the University of California at Berkeley. From 1991 to 1997, he was a research scientist with Schlumberger-Doll Research. From 1997 to 1999, he was a reservoir specialist and technology champion with Yacimientos Petrolíferos Fiscales in Buenos Aires, Argentina. Since 1999, he has been affiliated with the Department of Petroleum and Geosystems Engineering of The University of Texas at Austin, where he is currently a full professor, September 2011 SPE Journal 623

holds the Zarrow Centennial Professorship in Petroleum Engineering, and conducts research on borehole geophysics, formation evaluation, well logging, and integrated reservoir characterization. Torres-Verdín is the founder and director of the Research Consortium on Formation Evaluation at The University of Texas at Austin. He has served as a guest editor for Radio Science, the Journal of Electromagnetic Waves and Applications, and the SPE Journal and is an associate editor for Petrophysics and Geophysics. Torres-Verdín received the 2006 Distinguished Technical Achievement Award from the Society of Petrophysicists and Well Log Analysts (SPWLA); the 2008 Formation Evaluation Award from SPE; the 2003, 2004, 2006, and 2007 Best Paper Awards in Petrophysics by the SPWLA; and the 2006 Best Presentation Award and the 2007 Best Poster Award by SPWLA. Gong Li Wang holds a BS degree in geophysical well logging and a PhD degree in applied geophysics from the University of Petroleum, East China. From 1993 to 2002, he worked with the University of Petroleum, East China, on electromagnetic well logging. From 2002 to 2008, he worked with the Center for Computational Electromagnetics and the Electromagnetics Laboratory of the University of Illinois at Urbana-Champaign, and the Center for Petroleum and Geosystems Engineering of The University of Texas at Austin. Since 2008, he has been with Schlumberger as a tool physicist. In 2001, he was awarded the Second Prize of the Award for Technological Innovation by the China National Petroleum Corporation. His research interests are in wave propagation in inhomogeneous media, numerical modeling methods, and inverse problems. 624 September 2011 SPE Journal