spe-166178_geological attributes from conventional well logs

Upload: sebastian-ruiz-villa

Post on 24-Feb-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    1/15

    SPE 166178

    Geological Attributes from Conventional Well Logs: Relating Rock Types toDepositional Facies in Deepwater Turbidite ReservoirsChicheng Xu, SPE, The University of Texas at Austin; Carlos Torres-Verdn, SPE, The University of Texas at

    Austin; Ronald J. Steel, The University of Texas at Austin

    Copyright 2013, Society of Petroleum Engineers

    This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 30 September2 October 2013.

    This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not beenreviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, itsofficers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission toreproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

    Abstract

    Grain size and bed thickness are important geological attributes for sedimentary facies analysis and reservoir quality ranking

    in deepwater turbidite reservoirs. Grain size controls reservoir quality and has causative or correlational effects on most well

    logs. Additionally, bed thickness affects well logs in various ways because of different vertical logging-tool resolutions. Theobjective of this paper is to quantitatively classify rock and bed types based on conventional well logs to assist facies

    interpretation and stratigraphic reservoir modeling.

    We model physical properties and well-log responses originating from clastic reservoirs with different grain-size distributionsin the context of deepwater turbidite systems. The model accounts for fluid effects introduced by capillary transitions and

    salty connate water. Petrophysical relationships are examined between grain sizes and pore geometry as inferred from the

    combined effects of initial connate-water saturation and reservoir capillary pressure. From modeling, we derive several

    quantitative petrophysical attributes which correlate significantly with grain size. A list of such attributes includesvolumetric concentration of shale, total porosity, neutron-density porosity difference, and reservoir quality index (RQI). Rockclassification is then performed based on a relevant set of attributes to detect and rank different rock types with specific

    grain-size distributions. Concomitantly, a quantitative rock classification scheme using three-dimensional Thomas-Stieber

    diagrams constructed with gamma ray, bulk density, and resistivity logs is used to classify petrophysical zones based on theiraverage bed thicknesses. A dual-attribute rock classification scheme is proposed to better associate rock types with

    depositional facies.

    The above-mentioned rock-classification and bed-typing methods are synthesized and applied to a turbidite field from the

    Gulf of Mexico. We perform facies analysis based on the vertical succession of rock types (grain sizes) and the inferred

    variations of bed thickness. The field case demonstrates that these two geologic attributes have great potential in reducing

    uncertainty and non-uniqueness in the construction of stratigraphic reservoir models. Results indicate that the classified rock

    types and their petrophysical ranking agree with core data while the interpreted facies fit into the depositional context asinferred from core descriptions and seismic horizons.

    Introduction

    Rock classification (or simply rock typing) is an emerging reservoir description tool that elicits broad research interests in the

    oil and gas industry. It has become increasingly important in modern reservoir description to invoke close integration of

    multi-discipline, multi-physics, and multi-scale subsurface data including pore imaging, core measurements, well logs,seismic amplitude data, well testing, and production surveillance (Gunter et al., 1997a). Interestingly, the definition of rock

    type still remains ambiguous and sometimes arguable in many situations as it depends highly on the specific description

    purposes to be achieved by geoscientists and engineers. Geologists consider rock types as depositional facies or lithofacies

    which emphasize the genesis of rock formations to enable 3D stratigraphic reservoir modeling (Fisher, 1982; Kerans and

    Tinker, 1999; Muto and Steel, 2007; Slatt, 2007). Petrophysicists define rock types based on pore geometry that relates allstatic and dynamic petrophysical properties (Archie, 1950 and 1952). Reservoir and production engineers group rock types as

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    2/15

    2 SPE 166178

    flow units that are stratigraphically continuous intervals of similar geologic and petrophysical features to upscale reservoir

    grids for efficient fluid-flow simulation (Gunter et al., 1997b). Noteworthy is that reservoir characterization teams are

    working toward a common objective: to construct a verifiable reservoir model populated with accurate petrophysicalproperties for reserves estimation and production forecasting (Lucia, 1999). Therefore, developing new rock classification

    schemes and workflows that serve multiple characterization purposes from different disciplines remains a challenging but

    important task.

    For subsurface studies, whole core samples and well logs are the main sources of high-resolution geological data. Routinegeological work infers grain size and bed thickness from slabbed core samples. However, in most cases whole core spans a

    very limited segment of the reservoir. Log-core integration is therefore necessary to extrapolate geological properties fromcored depth intervals to remaining wells and into the reservoir. Rock classification based on core data is relatively simple

    because petrophysical properties are rigorously measured under the same protocol. However, rock classification using

    downhole well logs is more challenging and is subject to great uncertainty. Four technical issues immediately arise when

    moving from core data to well logs for rock typing, namely data quality, indirect measurements, variable reservoirconditions, and scale discrepancy (Xu, 2013). These compounded technical problems were initially defined by Archie as the

    geometrical problem (1950). In the present paper, we aim to address the issue of indirectness. Indirectness lies in using

    physical measurements to estimate petrophysical properties and interpret geological attributes. In fact, most well logs are

    simultaneously sensitive to grains, texture, structure, and fluids. Table 1.1 summarizes the sensitivity of various well logs to

    rock petrophysical properties and geological attributes. It is found that none of the logs are directly sensitive to grain size andonly a few logs have good sensitivity to bedding structure. Grain size information can only be indirectly inferred from the

    pore geometry formed in-between grains. Moreover, well logs highly sensitive to pore geometry and bedding structure arealso significantly influenced by fluid content, as in the case of resistivity logs. Therefore, fluid effects on well logs need to be

    cautiously accounted for before inferring grain and bed properties.

    Table 1Sensitivity of well logs to petrophysical properties and geological attributes (modifiedfrom Serra and Abbot, 1980).

    Composition Texture Structure Fluid

    Gamma Ray High Low Low Low

    PE Factor High Low Low Low

    Neutron Porosity Medium Medium Low High

    Bulk Density Medium Medium Low High

    Electrical Resistivity Low Medium High High

    Acoustic Slowness High Medium Medium High

    Magnetic Resonance Medium High Low High

    The ultimate goal of rock classification is to associate petrophysical rock types with geological framework and reconcilethem with seismic facies to propagate rock-type associated petrophysical properties into the reservoir model. This endeavor is

    difficult without understanding the cause-and-effect relationship between pore geometry and rock geological attributes such

    as grain size and bed thickness. Hence, it is necessary to investigate the sensitivity of well logs to those geological attributes

    and the associated pore geometries in the presence of different fluid-saturation conditions resulting from either hydrocarbon

    migration or mud-filtrate invasion. In this paper, we show the causative and correlational impacts of grain size on various

    types of logs. In addition, the relationships between gamma ray log and other logs are interpreted to reveal the underlying bedthickness information as emphasized in Thomas-Stieber diagrams. We propose an integrated petrophysical workflow and a

    dual-attribute rock classification scheme to better relate rock types to geological facies in deepwater turbidite reservoirs. A

    field case from the central Gulf of Mexico is used to test the workflow and classification schemes.

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    3/15

    SPE 166178 3

    Grain-Size Analysis from Well Logs

    Grain size is directly associated with depositional energy. Large grains are deposited under highly energetic flow conditions,whereas small grains are deposited under low flow energy (Reading, 1996). Grain size information provides key data for

    geologists in their facies interpretation work (Glaister and Nelson, 1974). In addition, grain-size distribution determines the

    pore-size distribution of most clastic rocks when diagenesis effects are not significant. Consequently, grain size information

    becomes a good indicator of reservoir quality in terms of hydraulic capacity.

    Grain size is typically measured from core samples or outcrops with a laser particle size analyzer (LPSA). To date, there

    exists no effective physical measurement capable of measuring grain size directly at downhole conditions. However, there aresome causative and correlational effects between grain size and physical logs. In the following section, we briefly review the

    relationships between grain size and gamma ray, neutron, density, resistivity, and NMR logs.

    Gamma Ray Log. Gamma ray logs are sensitive to the natural radioactivity of rock formations. The radioactive elements (U,K, Th) are predominantly associated with small grains such as silt or clay (Serra and Serra, 2003). Consequently, often there

    is a correlation between gamma ray log and grain-size distribution. In fact, the gamma ray log has been long used as a single

    indicator of grain size in many geological field studies because the gamma ray log is more readily available than other logs

    (Pirson, 1983; Rider, 1990). In clastic reservoirs, the gamma ray log is sufficient to separate clean sands, shaly sands, and

    shales (or siltstones). Figure 4 Track 2 shows the typical gamma ray log responses of clean sands, shaly sands, and shales (orsiltstones). However, the gamma ray log alone cannot effectively distinguish clean sands of different grain sizes.

    Neutron Porosity and Bulk Density Log. Both neutron porosity and bulk density logs are dominantly sensitive to total pore

    volume and pore-filing fluids in clean sands. In many clastic reservoirs, the porosity formed by grain deposition has small

    variability. However, the total pore volume will be significantly changed by cementation which is mainly comprised of silts

    and clays (Neasham, 1977). As a result, low porosity rock tends to be associated with silt- or clay-size grains. In addition, the

    presence of silts or clays will increase the bound water volume that remains in the rock pore system after hydrocarbonmigration. The hydroxyl in clay minerals gives rise to a high hydrogen index which increases the apparent neutron porosity.

    On the other hand, clay or silt cementation increases the bulk density of rocks when solid cement replaces the original fluid in

    the pore space. As a result, neutron and density porosity logs separate when they are plotted on the same track using the

    correct matrix. The separation or difference between neutron and density porosity is an indicator of clay volume. Figure 4,Track 3 shows the typical neutron porosity and bulk density logs across clean sands, shaly sands, and shales. Neutron-density

    separation is smallest in clean sands and largest in shales. However, neutron porosity and bulk density logs still lack the

    ability to differentiate sandstones of different grain sizes but with the same porosity.

    Resistivity and NMR Logs.We use a simple pore network model to study the cause-effect chain connecting a series of rockproperties including grain size, pore size, capillary pressure, water saturation, electrical conductivity, and NMR transverse

    relaxation time T2. Figure 1 shows two thin sections of pore network systems formed by large (Type A) and small (Type B)

    grain sizes, respectively. Large grain sizes generate large pore sizes accordingly, resulting in a T2distribution with large T2values (Fig. 2). When these two rocks are subjected to the same capillary pressure induced by hydrocarbon migration at

    reservoir conditions, rock type A exhibits both lower capillary-bound water saturation and clay hydration water saturation

    (Hill et al., 1979), and consequently lower electrical conductivity (Fig. 3). Therefore, in a hydrocarbon-bearing zone,resistivity logs become sensitive to pore size, i.e., petrophysical rock types. Rock types with higher hydraulic conductivity

    tend to have lower electrical conductivity (or higher resistivity) in hydrocarbon-bearing zones in the presence of capillary

    pressure. Figure 4 Track 4 describes the typical resistivity log responses observed across clean sands, shaly sands, and shalesin a hydrocarbon zone.

    Figure 1Thin sections describing pore network systems formed by different grain sizes. Left Panel: large grain size; Right Panel:small grain size (Xu et al., 2013).

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    4/15

    4 SPE 166178

    Figure 2Modeling of pore-size (left panel) and NMR T2-distributions (right panel) for two different rock types (Xu et al., 2013).

    Figure 3Modeling of drainage capillary pressure and resistivity-height relation for two different rock types. Pc: capillary pressure(left panel); HAFWL: height above the free water level (FWL) (right panel); Rt: bulk rock resistivity (Xu et al., 2013).

    Figure 4Simulated well logs across different rock types. Track 1: Depth; Track 2: Gamma ray log; Track 3: Neutron porosity(sandstone porosity units) and bulk density logs; Track 4: Resistivity log; Track 5: Rock types. Logs are simulated with softwareUTAPWeLS

    1.

    1The University of Texas at Austins Petrophysics and Well Log Simulator

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    5/15

    SPE 166178 5

    Correction of Free Water in Capillary Transition Zones.The presence of free water in a capillary transition zone disrupts

    the correlation between electrical conductivity and hydraulic conductivity. Xu and Torres-Verdn (2012) proposed a

    correction method by linking Leveretts RQI with in-situ capillary pressure (Pc) and initial connate-water saturation (Sw)using the empiricalJ-function model (Darling, 2005; Torres-Verdn, 2012):

    1

    ( ) ( )b w wirr bw wirr wS S

    S S aJ J S a

    = + =

    , (1)

    together with Leveretts capillary pressure model (1941), given by:

    1( ) cos

    RQI cos ( )w w wirr b

    c c

    J S S Sk

    P a P

    = = =

    , (2)

    where Swis initial water saturation, Swirris irreducible water saturation (which is set to 0.01 lower than the minimum watersaturation in the entire reservoir column, Darling, 2005); aand bare constants derived from core-measured capillary pressure

    curves; cos is the product of interfacial tension and contact angle, which is assumed constant in the same reservoir; Pcis

    in-situ reservoir capillary pressure in psi.

    Shale Distribution and Bed Typing from Well Logs

    Bedding structure together with bed thickness is often a key parameter for geologists to interpret sedimentary facies

    (Campbell, 1967). Thinly bedded or laminated formations are persistent only in some special depositional facies (Passey etal., 2006). High-resolution image logs provide the best estimation of bedding structures but are not commonly available in all

    wells. Thomas-Stiebers diagram (Thomas and Stieber, 1975 & 1977) is a widely applied petrophysical interpretation toolused to diagnose shale distribution in siliciclastic reservoirs as being laminated, dispersed, or structural (Fig.5; Torres-Verdn,2012). Thomas-Stieber diagram assumes two end members: clean-sand and pure-shale. Several trends are also defined to

    reveal different geological processes yielding different shale distributions which are associated with different mixing laws

    reflected on logging measurement physics and the underlying petrophysical properties (Xu and Torres-Verdn, 2013). As a

    consequence, rocks with different values of shale volumetric concentration and distribution appear in different locations

    when projected onto a Thomas-Stieber diagram (Fig. 6). Xu and Torres-Verdn (2013) proposed a graphical clusteringmethod to classify petrophysical zones of different bed thickness in a 3D Thomas-Stieber diagram (Fig. 7).

    (a) (b) (c)

    Figure 5Three nominal forms of shale distributions in siliciclastic reservoirs: (a) laminated shale, (b) dispersed shale, and (c)structural shale.

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    6/15

    6 SPE 166178

    Figure 6Thomas-Stieber diagram constructed with bulk density and gamma ray logs. Red circle dots: pivoting points; Blue dashed

    lines: trend lines; shaded elliptical regions: uncertainty bounds.

    Figure 7Rock classification and bed typing via k-means cluster analysis (Press et al., 2007)on a 3D Thomas-Stieber diagramconstructed with gamma ray, bulk density, and resistivity logs.

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    7/15

    SPE 166178 7

    Figure 8Integrated dual-attribute rock classification workflow based on well logs and core data.

    Dual-Attribute Rock Classification Workflow

    Rock types defined by grain size and bed thickness should be integrated to achieve a better reservoir description in multi-well

    field studies,. Figure 8 shows an integrated dual-attribute rock classification workflow based on standard petrophysical log-analysis and core-calibration. Grain size and bed thickness information are propagated from cored intervals to log-covered

    intervals via step-by-step calibration and validation. Therefore, the final rock types derived from well logs in multiple wells

    are geologically consistent for facies interpretation and stratigraphic reservoir modeling. Furthermore, rock type information

    needs to be returned to standard petrophysical analysis for more accurate rock-type-specific processing.

    Field Case: Marco Polo Field, Gulf of Mexico

    The Marco Polo field is a Miocene turbidite oil field located in the central Gulf of Mexico (Fig. 9; Contreras et al., 2006).Marco Polo minibasin resides in a tectonically active salt environment (Fig. 10; Jackson et al, 2008). The depositional system

    is interpreted as a submarine fan complex developed in a mini-basin with stacked progradational turbidite lobes. Reservoirsconsist primarily of sandy turbidite facies interbedded with some siltstone and muddy debris-flow facies. Reservoir rocks are

    mainly unconsolidated sandstones of very fine to fine- to medium-size grains. Thin-bedded zones are common across the

    entire reservoir. In this study, we evaluate only the two hydraulically connected, hydrocarbon-bearing sandstone units: M40

    and M50 in three wells. These sands are buried at depths between 12,000 and 13,000 ft TVD and dip toward the west (Fig.11). Core measurements indicate that the structureless and plane-parallel laminated sandstones exhibit porosities as high as

    35%, as well as 100 2000 md of nominal permeability. Well logs from three wells are used for rock typing. Well no. 1 wasdrilled with oil-base mud and penetrated a capillary transition zone and the FWL. Wells nos. 2 and 3 were drilled with water-

    base mud where the M40-50 sands are located more than 500 ft above the FWL (Fig. 11).

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    8/15

    8 SPE 166178

    Figure 9Geographic location of the Marco Polo field in the Gulf of Mexico (Contreras, 2006).

    Figure 10Salt tectonics background of the Marco Polo minibasin (Jackson et al., 2008).

    Figure 11Seismic cross-section with the three well trajectories considered in this paper.

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    9/15

    SPE 166178 9

    Core-Based Rock Typing and Grain Size/Bed Thickness Interpretation

    Petrophysical rock types are first established using routine core porosity-permeability measurements. Six petrophysical rock

    types were identified from the histogram distribution of Leveretts RQI (Fig. 12). Table 2 summarizes the statistical

    distributions of petrophysical properties for each rock type. We observe that rock types 1 - 3 exhibit a largely overlappingporosity range. Their average horizontal permeability exhibits a ratio around 2 2.5, which indicates that their average pore-

    throat radius ratio is close to 1.5. Laser grain size measurements for rock types confirmed that the controlling factor for

    permeability was mean grain size. Reservoir quality decreases with decreasing grain size. Grain size measurements for rock

    types 2, 4, and 5 are listed in Table 3 for comparison. Different bed types identified from core photographs are compared inTable 4 for facies interpretation.

    Figure 12Core-based hydraulic rock typing via Leveretts RQI.

    Table 2 Statistics of porosity, permeability, and RQI for eachhydraulic rock type.

    Rock

    Type

    Porosity (p.u.) Permeability (md) RQI

    (m)

    RRT1 33.6 1.8 1099 291 56.8 7.2

    RRT2 33.3 2.1 549 130 40.4 4.9

    RRT3 31.5 4.4 179 63.7 23.4 3.4

    RRT4 24.1 6.1 42.9 19.1 13.1 2.1

    RRT5 19.7 3.9 11.3 3.9 7.5 1.1

    RRT6 19.6 3.7 4 1.5 4.4 0.86

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    10/15

    10 SPE 166178

    Table 3Comparison of grain-size distribution data for rock types 2, 4, and 5.

    Rock Type Grain Size Distribution Description Facies Interpretation

    2

    Dominated by sand-size

    grains with a narrow

    distribution in the large-

    size mode (i.e., goodsorting).

    Turbidite channel infill

    including channel lag.

    4

    Dominated by silt-sizegrains with only a small

    proportion of sand-sizegrains. Overall

    distribution is widely

    spread (i.e., relatively

    poor sorting).

    Channel levee or channelmargin with thin, fine-

    grained beds.

    5

    Dominated by silt-sizegrains with an overall

    wide distribution (i.e.,

    poor sorting).

    Overbank depositspossibly including thin

    debris flows.

    Table 4: Comparison of core photographs of different bed types.

    Rock Type Core Photos Description Facies Interpretation

    Thick sand

    Structureless coarse-grained

    sandstone with no clearinternal lamination.

    Amalgamated turbidite

    channel deposits.

    Laminated sand-shale

    Interbedded sand-shale

    sequence with variable net-to-gross ratio.

    Channel levee or margin

    deposits.

    Thick shale or siltstone

    Fine-grained shale or

    siltstone.

    Overbank deposits, possibly

    with debris flow.

    Sand Silt

    Sand

    Silt

    Sand Silt

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    11/15

    SPE 166178 11

    Log-Based Rock Classification and Bed Typing

    Core-calibrated, conventional well-log analysis is first performed to provide relevant petrophysical properties, such as

    volumetric concentration of shale, total porosity, and connate-water saturation. We classify rock types via k-means cluster

    analysis (Press et al., 2007) on several relevant petrophysical attributes including shale volume, total porosity, neutron-density separation, water saturation in the irreducible zone, and well-log-derived RQI as in Eq. (2) in the capillary transition

    zone. The classified rock types are consistent with core data concerning both petrophysical properties and grain size. Figure

    13 shows an interval with log-based rock classification results. A clear fining-upward grain-size trend is observed in both

    core and the log-derived rock types in the interval of 13,574 13,580 ft. The high-resolution grain-size information providesgeologists useful data for facies interpretation.

    Figure 13Example of a fining-upward unit identified from vertical stacking of petrophysical rock types. Track 1: Depth; Track 2:Gamma ray log; Track 3: Neutron (in apparent sandstone porosity units) porosity and bulk density logs; Track 4: Inductionresistivity logs; Track 5: Rock types of different grain-size distributions; Track 6: Core permeability.

    Relating Rock Types to Geological Facies in a Single Well

    In a single well, we analyze the stacking pattern or trend of the classified rock types that bear grain-size and bed-thickness

    information. Similar neighboring rock types are grouped as distinct facies from the adjacent rock type groups. Figures 14 and15 show two intervals with their sedimentary facies interpretations. Amalgamated channel facies are composed mainly of

    thick and coarse-grained sandstones. Both bed thickness and grain size decrease in the levee or channel margin facies.Overbank deposits are mainly shales or siltstones. The interpretation is consistent with observations contained in Table 3 and

    Table 4.

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    12/15

    12 SPE 166178

    Figure 14Rock typing and facies interpretation in the key well of the Gulf-of-Mexico case. Tracks 1-4: basic well logs; Tracks 5-6:standard well-log analysis, saturation-height for rock types 1 to 3; Tracks 7: distribution of rock types; Track 8: interpretedsedimentary facies.

    Figure 15Rock typing and facies interpretation in the key well of the Gulf-of-Mexico case. Tracks 1-4: basic well logs; Tracks 5-6:standard well-log analysis, saturation-height for rock types 1 to 3; Tracks 7: distribution of rock types; Track 8: interpretedsedimentary facies.

    8

    8

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    13/15

    SPE 166178 13

    Stratigraphic Correlation with Multi-Well Logs. After rock classification, bed typing, and facies interpretations are

    performed in individual wells, a multi-well cross-sectional view can be constructed for stratigraphic analysis. Figure 16

    shows the corresponding results from the three well penetrations. The locations of these wells in a submarine fan complexwith stacked progradational lobes are consistent with the interpretation made from seismic horizons. Stratigraphic analysis

    together with the estimated petrophysical properties can then be integrated with seismic amplitude data for populating a

    reservoir model.

    Figure 16Use of rock-type and bed-type distributions in multiple wells to assist sedimentological and stratigraphic interpretation.The upper panel shows the Thomas-Stieber diagram constructed with bulk density and gamma ray logs in each well below. Thelower panel shows the log tracks in each well with the following information: Track 1: Depth; Track 2: Gamma ray log; Track 3:Neutron (in apparent sandstone porosity units) porosity and bulk density logs; Track 4: Induction resistivity logs; Track 5: Bed type;Track 6: Rock type of different grain-size distribution.

    ConclusionsGrain size and bed thickness are two geological attributes that can be interpreted from conventional logs in turbidite

    reservoirs. The petrophysical quality of young deepwater turbidite reservoirs is controlled chiefly by sedimentary grain sizes.

    Therefore, a link exists between petrophysical rock types and depositional facies which can be effectively applied in reservoir

    development efforts. Thomas-Stieber diagrams are useful in identifying thinly bedded or laminated zones which provide

    additional aid for facies interpretation. We propose an integrated petrophysical workflow that combines grain-size and bed-thickness attributes to reduce uncertainty in stratigraphic reservoir modeling. Reservoir capillary transition effects on

    resistivity logs need to be corrected for more accurate estimation of grain sizes. In the deepwater Gulf-of-Mexico field case,

    we confirmed that multi-well rock typing can assist sedimentary and stratigraphic studies which have potential applicationsfor exploration in frontier basins or in deepwater environments, where subsalt or pre-salt reservoirs cannot be clearly imaged

    with seismic data and well drilling is limited due to high rig costs.

    30 40 50 60 70 80 90 100

    1.8

    2

    2.2

    2.4

    2.6

    2 2

    2886

    255

    55 2255

    2555 55556

    65

    5555

    55

    22 22

    2566666

    4 1 665

    50

    50 7

    75

    0

    0500 53

    7 7

    7737

    7755 7

    3

    75

    5000 0 5

    5 5 7556

    66 6

    6 55

    9

    7

    5

    555 5

    22222 222 2222

    222225 2 2225

    222

    222 222 222

    2

    5

    54

    6

    2 2 2222 222 2 222

    26

    66

    5

    7

    599 9

    77

    97777 7

    7

    7

    9 9

    977777 9

    7

    7

    77

    75 7

    7 7

    77

    77

    77

    77

    9 99

    99

    999

    99997

    3 3

    7

    7

    7

    3

    00

    0 0 77 7

    333

    00

    0000

    0000

    0 00 0

    000 70

    0 3 7

    5 777

    75

    733 300 335000

    5

    5

    6

    5

    500 0

    000 5555 77

    7

    00 330

    7

    5 97

    7

    5

    9

    9

    77

    5

    7

    777

    505 5 7

    9

    9

    77

    9 9

    25

    77

    73

    5 500000

    05

    5 5 55

    55

    5555

    56

    6 55

    2 2222 2222222 22222

    222 2222

    2 22 2 22222

    GR (GAPI)

    RHOB(G/CM3)

    30 40 50 60 70 80 90 100

    1.8

    2

    2.2

    2.4

    2.6

    2.8

    222

    2226 2

    2255 25 525

    5 2 22225 22 22666 666

    7

    773330000073 70030

    0 7 7

    773333

    377

    57

    337

    3300

    000 000

    3 333 370 777

    05

    770 7

    7

    5

    6 222

    55

    5 595

    755

    5

    5

    62 2222 2 2

    2 2222 22222 2222 222

    2

    22

    2222 2222

    2 2222228

    5

    55777 7

    7 9 99

    9 9977 9 997

    7

    77

    7 99

    99977 997

    79

    97

    977 7

    7 9

    7 7

    779

    999 99 999

    99999

    979

    77775

    775 7

    73 7 777333300 0050 50 5055 557

    50 7

    7

    7703 37

    5

    5

    6 5 2 22

    2 2222 2 22

    2222222

    2222

    22 22 22 222 2

    GR (GAPI)

    RHOB(G/CM3)

    30 40 50 60 70 80 90 100

    1.8

    2

    2.2

    2.4

    2.6

    2.8

    2228 22 2 22

    22

    22 22 2

    222 222 226

    2822

    55 2 22 2222 22 2 222

    222 22

    2 2

    29

    5

    77

    000 00000 5

    500000000 5 7700

    30 700 05 5

    7705 5

    550 500 5 55500 550 5 55555 9

    2

    5

    2222 2

    2

    2255

    29

    9

    6

    6 6 6 5 2 222

    2222

    55

    5 5

    66

    5 9995 5555555

    56 5

    97

    6 5 5555

    9555 5

    55

    9

    97

    5

    5

    555

    99

    6

    1 6 5

    29

    97 775

    5 5555500

    65

    97775 5 977

    55

    775555 555000 550 55

    5555505 7

    555 7

    9997 9

    977 7 999

    599

    9 95

    9955

    5 22 2

    22 22 2

    -12

    22

    222

    2222

    222 22 2

    -1

    -1

    9

    7

    GR (GAPI)

    RHOB

    (G/CM3)

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    14/15

    14 SPE 166178

    ACKNOWLEDGMENTSWe would like to thank Anadarko Petroleum Cooperation for providing the field data used for analysis and verification. The

    work reported in this paper was funded by The University of Texas at Austins Research Consortium on FormationEvaluation, jointly sponsored by Afren, Anadarko, Apache, Aramco, Baker-Hughes, BG, BHP Billiton, BP, Chevron, China

    Oilfield Services, LTD., ConocoPhillips, ENI, ExxonMobil, Halliburton, Hess, Maersk, Marathon Oil Corporation, Mexican

    Institute for Petroleum, Nexen, ONGC, OXY, Petrobras, PTTEP, Repsol, RWE, Schlumberger, Shell, Statoil, Total,

    Weatherford, Wintershall and Woodside Petroleum Limited.

    Nomenclaturea : Linear coefficient in empiricalJfunction, []b : Exponent in empiricalJfunction, []

    J : LeverettsJfunction, []

    k : Permeability, [mD]p : Density function of pore-throat size, []Pc : Capillary pressure, [psi]

    R : Pore-throat radius, [m]

    Rt

    : Bulk rock resistivity, [ohm.m]

    Rw : Connate water resistivity, [ohm.m]Sw : Water saturation, [frac]Swirr : Irreducible water saturation, [frac]

    T2 : NMR Transverse relaxation time, [ms]

    : Total porosity, [frac]

    : Interfacial tension, [dyne/cm] : Contact angle, [degree]

    List of Acronyms and Log Curve MnemonicsAHT1-9 : Array Induction Log (H) with Two Feet Resolution and 10-90 inch Depth of Investigation

    BS : Bit Size LogDRHO : Density Correction Log

    FWL : Free Water Level

    GR : Gamma Ray Log

    HAFWL : Height Above Free Water Level

    HCAL : Caliper LogHGR : Gamma Ray Log (High Resolution)

    K : Potassium

    K_Klink : Permeability Corrected with Klinkenberg EffectLPSA : Laser Particle Size Analyzer

    MICP : Mercury Injection Capillary Pressure

    NMR : Nuclear Magnetic Resonance

    NPHI : Neutron Porosity LogPcow : Capillary Pressure between Oil and Water

    PE : Photo-Electric

    PHI_LOG : Porosity from Log

    Por : Core Porosity

    RHOB : Bulk Density LogRHOZ : Bulk Density Log (High Resolution)

    RQI : Reservoir Quality Index

    RRT : Reservoir Rock TypeRT : Rock Type

    SH : Shale

    SS : SandstoneSw_c : Core Water Saturation

    Sw_RT : Water Saturation for Rock Type

    Swt_LOG : Water Saturation from Logs

    T2LM : Algorithmic Mean of NMR Transverse relaxation timeT2

    Th : ThoriumTVD : Total Vertical Depth

    U : Uranium

    UTAPWeLS : The University of Texas at Austins Petrophysics and Well Log Simulator

  • 7/25/2019 SPE-166178_Geological Attributes From Conventional Well Logs

    15/15

    SPE 166178 15

    References

    Archie, G.E. 1950. Introduction to Petrophysics of Reservoir Rocks.AAPG Bulletin, 34(5): 943 - 961.Archie, G.E. 1952. Classification of Carbonate Reservoir Rocks and Petrophysical Considerations.AAPG Bulletin, 36(2): 278-298.Campbell, C. V. 1967. Lamiae, Laminaset, Bed and Bedset. Sedimentology, 8(1967): 7-26.Contreras A. 2006. Spatial Delineation, Fluid-Lithology Characterization, and Petrophysical Modeling of Deepwater Gulf of Mexico

    Reservoirs through Joint AVA Deterministic and Stochastic Inversion of 3D Partially-Stacked Seismic Amplitude Data and WellLogs. Ph. D. Dissertation, The University of Texas at Austin.

    Contreras A., Torres-Verdn, C., Chesters, W., Kvien, K., and Fasnacht T. 2006. Extrapolation of Flow Units away from Wells with 3DPre-Stack Seismic Amplitude Data: Field Example. Petrophysics, 47(3): 223-238.

    Darling, T. 2005. Well Logging and Formation Evaluation. Gulf Professional Publishing.Fisher, W. L. 1982. Depositional Systems and Environments, in Encyclopedia of Science and Technology. New York, McGraw-Hill, 5th

    ed., 104-108.Glaister, R.P., and Nelson, H.W. 1974. Grains-Size Distributions, an Aid in Facies Identification.Bulletin of Canadian Petroleum Geology,

    22(3):203-240.Gunter, G.W., Pinch, J.J., Finneran, J.M., and Bryant, W.T. 1997a. Overview of an Integrated Process Model to Develop Petrophysical

    Based Reservoir Descriptions. Paper SPE 38748 presented at the SPE Annual Technical Conference and Exhibition, San Antonio,Texas, 5-8 October.

    Gunter, G.W., Finneran, J.M., Hartmann, D.J., and Miller, J.D. 1997b. Early Determination of Reservoir Flow Units Using an IntegratedPetrophysical Method. Paper SPE 38679 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, 5-8

    October.Hill, H.J., Shirley, O.J., and Klein, G.E. 1979. Bound Water in Shaly Sands Its Relation to Qv and Other Formation parameters. The Log

    Analyst, May-June, 3 -19.

    Jackson, M.P.A., Hudec, M.R., and Dooley, T.P. 2010. Some Emerging Concepts in Salt Tectonics in the Deepwater Gulf of Mexico:Intrusive Plumes, Canopy-Margin Thrusts, Minibasin Triggers and Allochthonous Fragments. Vining, B.A. and Pickering, S.C. (eds)Petroleum Geology: From Mature Basins to New Frontiers Proceedings of the 7th Petroleum Geology Conference, 899912.

    Kerans, C., and Tinker, S. 1999. Sequence Stratigraphy and Characterization of Carbonate Reservoirs. SEPM short course no. 40.Leverett, M.C. 1941. Capillary Behavior in Porous Solids. Transactions of the AIME, 142(1): 159172.

    Lucia, F.J. 1999. Carbonate reservoir characterization. Springer-Verlag, Berlin.Muto, T., Steel, R., and Swenson, J. 2007. Autostratigraphy: a Framework norm for Genetic Stratigraphy. Journal of Sedimentary

    Research, 77: 2-12.

    Neasham, J.W. 1977. The Morphology of Dispersed Clay in Sandstone Reservoirs and Its Effect on Sandstone Shaliness, Pore Space, andFluid Flow Properties. Paper SPE-6858 presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 9-12October.

    Passey, Q.R., Dahlberg, K.E., Sullivan, K., Yin, H., Brackett, B., Xiao, Y., and Guzman-Garcia, A.G. 2006. Petrophysical Evaluation ofHydrocarbon Pore-Thickness in Thinly Bedded Clastic Reservoirs. AAPG Publications, Archie Series, No.1.

    Pirson J.S. 1983. Geologic Well Log Analysis. Third Edition, Gulf Professional Publishing.Press, W.H., Teukolsky, S.A., Vetterling, W.T., and Flannery, B.P. 2007. Section 16.1: Gaussian Mixture Models and k-Means Clustering,

    Numerical Recipes. The Art of Scientific Computing, 3rd Ed. New York: Cambridge University Press.

    Reading, H.G. 1996. Sedimentary Environments: Processes, Facies and Stratigraphy. 3rd

    Edition, Blackwell Science.Rider, M.H. 1990. Gamma-Ray Log Shape Used as a Facies Indicator: Critical Analysis of an Oversimplified Methodology. Geological

    Society, London, Special publications, 48: (27-37).Serra, O., and Abbott, H. T. 1980. The Contribution of Logging Data to Sedimentology and Stratigraphy. Paper SPE 9270 presented at 55th

    SPE Annual Fall Technical Conference and Exhibition, Dallas, Texas, 21-24 September.Serra, O., and Serra, L. 2003. Well Logging and Geology. Technip.Slatt, R.M. 2007. Stratigraphic Reservoir Characterization for Petroleum Geologists, Geophysicists, and Engineers . Elsevier Publication.Thomas, E. C., and Stieber, S. J. 1975. The Distribution of Shale in Sandstones and Its Effect Upon Porosity. Paper T presented at the 16th

    Annual Logging Symposium of the SPWLA, New Orleans, Louisiana, 4-7 June.Thomas, E. C., and Stieber, S. J. 1977. Log Derived Shale Distributions in Sandstone and Its Effect upon Porosity, Water Saturation, and

    Permeability. 6th Formation Evaluation Symposium of the Canadian Well Logging Society, Calgary, Alberta, 24-26 October.Torres-Verdn, C. 2012. Integrated Geological-Petrophysical Interpretation of Well Logs. Lecture Notes for Fundamental Well Logging

    Course in the University of Texas at Austin, Unpublished.

    Xu, C. 2013. Reservoir Description by Well-Log-Based and Core-Calibrated Petrophysical Rock Classification. Ph. D. Dissertation, TheUniversity of Texas at Austin.

    Xu, C., and Torres-Verdn, C. 2012. Saturation-Height and Invasion Consistent Hydraulic Rock Typing Using Multi-Well ConventionalLogs. Paper KK presented at SPWLA 53rd Annual Logging Symposium, Cartagena, Colombia, 16-20 June.

    Xu, C., Torres-Verdn, C., and Gao, S. 2013. Electrical versus Hydraulic Rock Types in Clastic Reservoirs: Pore-Scale Modeling Verified

    with Field Observations in the Gulf of Mexico, U.S. 83rd

    SEG Annual Meeting, Houston, 22 27 September.Xu, C., and Torres-Verdn, C. 2013.Quantitative Rock Type Codification and Interpretation Using Log-Attributes Derived from Thomas-

    Stieber Diagrams in Siliciclastic Reservoirs. SEG Interpretation Journal (submitted for review).