You are on page 1of 9

Integrated Reservoir Characterization

Study of a Carbonate Ramp Reservoir:


Seminole San Andres Unit, Gaines
County, Texas
F.P. Wang,* SPE, F. Jerry Lucia, SPE, and Charles Kerans, SPE, Bureau of Economic Geology,
The U. of Texas at Austin

Summary

One of the important issues in constructing geologic and reservoir


models is to define geologic frameworks. A geologic framework is
fundamental to defining flow units, to interpolating well data, and
thereby to modeling fluid flow. For the Seminole San Andres Unit
(SSAU), the high-frequency cycles (HFCs) and rock-fabric facies
identified on outcrop analogs and cores were used to correlate
wireline logs. Reservoir and simulation models of the outcrop and
a two-section area of SSAU were constructed with rock-fabric units
within the HFCs as a geologic framework. Simulations were
performed using these models to investigate critical factors affecting recovery.
HFCs and rock-fabric units are the two critical scales for
modeling shallow-water carbonate ramp reservoirs. Descriptions of
rock-fabric facies stacked within HFCs provide the most accurate
framework for constructing geologic and reservoir models, because
discrete petrophysical functions can be fit to rock fabrics and fluid
flow can be approximated by the kh ratios among rock-fabric flow
units. Permeability is calculated using rock-fabric-specific transforms between interparticle porosity and permeability. Core analysis data showed that separate-vug porosity has a very strong effect
on relative permeability and capillary pressure measurements.
The effect of stratigraphic constraints on stochastic simulation
was studied. Geologic models generated by a conventional linear
interpolation, a stochastic simulation with stratigraphic constraints,
and a stochastic simulation without stratigraphic constraints were
compared. The stratigraphic features of carbonates can be observed
in stochastic realizations only when they are constrained by rockfabric flow units. Simulation results from these realizations are
similar in recovery but different in production and injection rates.
Scale-up of permeability in the vertical direction was investigated in terms of the ratio of vertical permeability to horizontal
permeability (k vh ). This ratio decreases exponentially with the
vertical gridblock size up to the average cycle size of 20 ft (6.1 m)
and remains at a value of 0.06 for a gridblock size of more than 20
ft (.6.1 m), which is the average thickness of HFCs. Simulation
results showed that the critical factors affecting recovery efficiency
are stacking patterns of rock-fabric flow units, k vh ratio, and dense
mudstone distribution.
Introduction

The SSAU lies on the northeastern margin of Central Basin Platform (Fig. 1) immediately south of the San Simon Channel.1 It
covers approximately 23 sq miles and contains more than 600 wells.
The field, discovered in 1936, is a solution-gas-drive reservoir with
a small initial gas cap, and it has an estimated original oil in place
(OOIP) of 1,100 MMSTB.2 Production comes from the Upper San
Andres Formation and the upper part of the Lower San Andres
Formation. The crude is 35API and has an initial formation volume
factor (FVF) of 1.39 and a solution/gas ratio of 684 scf/STB.
* Now with PGS Reservoir.
Copyright 1998 Society of Petroleum Engineers
Original SPE manuscript received for review 8 October 1996. Revised manuscript
received 10 November 1997. Paper peer approved 19 December 1998. Paper (SPE
36515) first presented at the 1996 SPE Annual Technical Conference and Exhibition,
Denver, 69 October.

SPE Reservoir Evaluation & Engineering, April 1998

The field was developed during the 1940s and produced 120
MMSTB (about 11% of OOIP) during the primary recovery from
1936 to 1969, in which time the reservoir pressure dropped from
2,020 to about 1,100 psig. Waterflooding was initiated in late 1969
using alternating rows of 160-acre inverted nine-spot patterns. Infill
drilling occurred in 1976, converting the pattern to a mixed 80- and
160-acre inverted nine spot. Waterflooding increased oil recovery
to 388 MMSTB. The characteristics of waterflooding are a short
fill-up time, a sharp increase in pressure, and a sharp decrease in
gas/oil ratio. A second infill drilling program that converted the
pattern to an 80-acre inverted nine spot occurred during 1984
through 1985. Fieldwide CO2 flooding began in 1985. The CO2
flooding further increased oil production, and the cumulative oil
production was about 539 MMSTB in 1994.
The SSAU has an excellent suite of cores and a large amount of
core, wireline log, and production data. A two-section area, Tract
2328, which has 33 wells with complete porosity log suites and
11 cored wells covering nearly the entire reservoir interval, was
selected for detailed geologic, petrophysical, and engineering characterization. This paper summarizes the results of this integrated
outcrop and subsurface characterization. More complete studies
were reported by Senger et al.,3 Kerans et al.,1 Wang et al.,4 and
Lucia et al.5 The objectives of this study were (1) to define critical
scales for constructing reservoir and simulation models for carbonate ramp reservoirs, (2) to study the effects of rock fabrics on
petrophysical properties, (3) to determine important geostatistical
parameters from outcrop and subsurface data, (4) to investigate the
effect of stratigraphic constraints on stochastic simulation and
recovery efficiency, and (5) to study factors affecting recovery
efficiency, such as the stacking patterns of rock-fabric units and k vh
ratio using outcrop and subsurface models.
Rock-Fabric and Petrophysical-Property
Relationships

Petrophysical properties of porosity, permeability, and saturation


are a function of pore-size distribution, which is related directly to
rock fabrics.6 The Seminole San Andres reservoir produces from
anhydritic vuggy and nonvuggy dolomites and contains three
rock-fabric/petrophysical classes7: (1) dolograinstone; (2) graindominated packstone and medium-crystalline mud-dominated dolostone; and (3) fine-crystalline mud-dominated dolostone (Table
1). These fabrics have unique stratigraphic locations and petrophysical characteristics.
Core Porosity and Core Permeability. The effect of rock fabrics
on porosity-permeability transforms was shown by Lucia6 and
subsequently was modified by Lucia et al.5 Data from the SSAU
2505 well and Lawyer Canyon outcrop were used to develop
porosity-permeability transforms for the SSAU. Rock fabrics and
pore types were described by point counting thin sections. Fig. 2
shows core porosity and permeability from the SSAU 2505 well.
For a given porosity, permeability varies with rock fabric by one
to two orders of magnitudes. Within a specific rock fabric, the
permeability increases with interparticle porosity.
Alternatively, these porosity-permeability transforms can be
used to estimate separate-vug porosity,4 because separate-vug porosity contributes little to permeability. Normally, separate-vug
105

Fig. 1Location map of Seminole San Andres Unit (SSAU) in the


Permian Basin, West Texas, and two section area (bold outline).

TABLE 1ROCK-FABRIC/PETROPHYSICAL CLASSES


COMMON IN THE SEMINOLE SAN ANDRES STUDY
Class

Rock Fabric

Particle Size (PS)

I
II

Dolograinstone
Grain-dominated dolopackstone
and medium-crystalline
dolowackestone
Mud-dominated dolopackstone
and fine-crystalline
dolowackestone

$100 mm
20 , PS , 100 mm

III

#20 mm

Fig. 2Porosity, permeability, and rock-fabric relationships for


grainstone, dolopackstone, and dolowackestone having crystal
sizes of 20 mm or less. Data from SSAU 2505 well.

porosity can be point counted from thin sections or estimated from


core slab surfaces. Because thin sections are often too small to
represent whole cores or sometimes are unavailable (which is
commonly the case with relative permeability data), separate-vug
porosity can be estimated by subtracting interparticle porosity from
total porosity where total porosity is the core porosity and interparticle porosity is obtained from the porosity-permeability transform for a specific rock fabric at a given permeability value.
Effect of Separate-Vug Porosity on Relative Permeability. The
effect of separate-vug porosity on capillary pressure and relative
permeability has not been studied extensively. Wang et al.,4 using
laboratory data measured on Lawyer Canyon outcrop and SSAU
core samples, showed that separate-vug porosity has a strong effect
on waterflood recovery and residual oil saturation. With additional
SSAU data, the correlation between waterflood recovery and separate-vug porosity is improved.
Although special core analyses were performed in 1981 on core
plugs from two wellsSSAU 2310 and 4902these core plugs
were not available for this study in 1992. Whole cores were
therefore sampled immediately adjacent to locations of the original
core plugs used for special core analyses, and thin sections were
made and rock fabrics were determined.
Three sets of relative permeability curves (Fig. 3a) measured
from SSAU 2310 cores show that the relative permeability increases with core permeability. Without examining the cores, the
low relative permeabilities of oil for samples 4WC and 9WC would
probably be explained as mixed and strongly oil wet. However, the
core porosity of 19.5% and a permeability of 5.6 md for sample
9WC indicate a high separate-vug porosity; the separate-vug porosity estimated from rock fabric, permeability, and porosity data
is 9%. The photomicrograph from thin sections adjacent to core
plug 9WC (Fig. 3b) clearly shows many separate intrafusulinid
106

Fig. 3(a) Three relative permeability curves from SSAU 2310,


and (b) photomicrograph of thin section at 5,229 ft adjacent to
core plug 9WC, showing both intrafusulinid (separate-vug) and
interparticle pores.
SPE Reservoir Evaluation & Engineering, April 1998

vugs and interparticle pores. The core data and the photomicrograph thus suggest that the separate-vug porosity is another factor
other than wettability that is controlling the relative permeabilities.
All the relative permeability data from SSAU 2310 and 4902
were summarized in terms of recovery and residual oil saturation
(Fig. 4a). Recovery by waterflooding decreases with an increase in
the ratio of separate-vug porosity to total porosity (vug porosity
ratio, or R vp ), whereas residual oil saturation increases with R vp .
Residual oil saturations determined from steady-state experiments
are much lower than those determined from unsteady-state experiments (Fig. 4b), because the real residual oil saturations were not
reached in the unsteady-state method.
Capillary Pressures. Capillary pressure data from the SSAU 2310
well are separated into grain-dominated dolopackstone and medium-crystalline mud-dominated dolostone (Fig. 5). In both rock
fabrics the capillary pressure decreases with an increase of porosity
but not of permeability. For example, vuggy carbonates have low
permeability and high porosity, but relatively low capillary pressure.
Reservoir Modeling

One of the important issues in constructing geologic and reservoir


models is to define geologic frameworks. A geologic framework is
fundamental to defining flow units, interpolating well data into
interwell regions, and thereby modeling fluid flow. Scales of
reservoir and flow simulation models vary considerably. However,
simulations using too coarsely scaled models are not representative,
whereas simulations using fine-scaled models are costly. Critical
scales are the scales at which depositional facies can be properly
correlated and petrophysical properties and fluid flow can be
properly modeled. For carbonates, two critical scales are HFCs and
rock-fabric units.1,8,9 The stacking of rock-fabric units in an HFC
defines the framework.

Fig. 4 (a) Waterflooding recovery and (b) residual oil saturation


as a function of vuggy porosity ratio (Rvp).
SPE Reservoir Evaluation & Engineering, April 1998

Fig. 5Capillary pressure, water saturation, and rock-fabric


relationships for medium-crystalline mud-dominated dolowackestone. Data from SSAU 2310 well.

The reason for using rock-fabric/petrophysical classes to define


flow units is that many petrophysical properties and correlations,
such as porosity-permeability transforms, capillary pressure, relative permeability, and residual oil saturation, can be better grouped
according to rock-fabric/petrophysical classes than to strict depositional facies groupings.7 For example, fusulinid packstone, ooidpeloid packstone, and dasycladacean-mollusk wackestone may all
behave petrophysically as mud-dominated rock-fabric facies and
therefore can be grouped into a single rock-fabric/petrophysical class.
Geologic Framework. A detailed reservoir characterization study
was carried out on a two-section area, Tract 2328 of the SSAU.
The depositional model used is of a carbonate ramp, a simple 0.2
to 2 seaward-sloping depositional interface that extends from just
above sea level to a depth of several hundred feet. Discrete
environmental belts on this ramp from landward to seaward include
inner ramp, ramp crest, and outer ramp. The ramp crest is the
critical belt where the fair-weather wave base intersects the depositional profile, creating a 1- to 2-mile-wide belt of higher energy,
grain-dominated rock-fabric facies. This belt, which occupies a
water-depth range of 0 to 30 ft, separates lower energy 0- to 10-ft
water-depth inner-ramp deposits landward from outer-ramp deposits seaward. The outer-ramp environment extends from 30 to 200
ft of water depth and is characterized by heterolithic mud-dominated rock-fabric facies.
Eleven cores covering the reservoir were described in detail.
Twelve HFCs and flow units as shown in Fig. 6 of the Amerada
Hess SSAU 2505 well were identified in cores and uncored wells.
The upper 9 HFCs (Cycles 1 to 9) record progradation of the
ramp-crest facies tract over the outer ramp during lower San Andres
composite sequence progradation. These HFCs are typical upward-shallowing cycles having basal mudstones and wackestones
grading upward into grain-dominated packstones and grainstones.
Rock-fabric variability includes thin intercalation of Classes I, II,
and III rock fabrics.7 The lower producing interval (Cycles 10 to
12) is composed of outer-ramp facies. The HFCs are composed of
dolowackestone and grain-dominated packstone fabrics. However,
pore size in these wackestone facies has been significantly enhanced by dolomitization, and thus rock fabrics fall largely within
petrophysical Class II.7
Petrophysics and Reservoir Model. To construct the reservoir
model, core data were calibrated with log data using neutron,
density, acoustic, and resistivity logs. Total porosity was calculated
using the neutron, density, and acoustic logs. Separatevug porosity
was calculated using a calibration of separate-vug porosity to total
porosity and acoustic transit time9 (Fig. 7). A Z-plot of total
porosity, water saturation, and rock fabric was used to define
rock-fabric fields. Petrophysical properties of total porosity, separate-vug porosity, water saturation, permeability, and rock fabrics
were calculated for 33 wells, and the results of the petrophysical
evaluation of the SSAU 2309 well are shown in Fig. 8. The
107

Fig. 7Wireline log/separate-vug porosity and rock-fabric relationships. Relationship between acoustic transit time and separate-vug porosity form thin-section point counts.

one simulation. This method is easy and fast, but the realizations
are too random to accurately depict realistic stratigraphic distributions. One of the recent trends in stochastic simulation is to generate
geologically realistic models using statistical techniques and engineering data. In several examples we found that deterministic
stratigraphic constraints are the most applicable.
Two stratigraphic constraints used are the rock-fabric flow units
and HFCs. Stochastic simulations were performed separately for
each rock-fabric unit using rock-fabric-specific geostatistical parameters. The realization for the entire reservoir is accomplished by
combining all realizations of individual rock-fabric units. Fig. 11
compares permeability distributions in Cycles 9 to 11 along a cross
section on the SSAU 2309 well, Tract 23-28. This comparison
includes examples generated by a conventional linear interpolation
and by stochastic simulations with and without stratigraphic constraints. The linearly interpolated permeability patterns are smooth
and continuous (Fig. 11a); the stochastically generated permeability
data without stratigraphic constraints (Fig. 11b) are too random, and
Fig. 6 Twelve high-frequency cycles and rock-fabric facies in
Amerada Hess SSAU 2505 well.

calculated permeability matches the core permeability better in


Cycles 6 and 8a than in Cycle 5 because of the differences in rock
fabric calculated from logs and from core data.
A three-dimensional (3D) reservoir model of the two-section
study area was constructed using a 3D geocellular modeling software. In modeling, the geologic framework was first built by
mapping the tops of 12 HFCs. Porosity, permeability, and water
saturation values calculated from petrophysical analyses of each
well location were interpolated among wells. At each location, the
vertical block sizes are the same within each cycle but are different
among cycles. The permeability distribution in a west-to-east cross
section shows that the permeability is generally more uniform and
higher in Cycles 9 to 12 than in Cycles 1 to 8 (Fig. 9). The 3D
porosity distributions in Cycles 8 and 9 (Fig. 10) show the upwardshallowing patterns and significant lateral variability within a
rock-fabric flow unit.
Because most reservoir simulation programs do not allow for
discontinuous layers, all flow-unit boundaries must be continuous
within the model. This results in rock-fabric flow layers containing
more than one rock-fabric facies. No sharp boundaries are placed
between the facies because no sharp boundaries have been found
in analog outcrops, and the average petrophysical values are interpolated between wells to fill the HFC framework.
Stochastic Simulation

One method commonly used in stochastic simulation is the generation of stochastically distributed data over the entire reservoir in
108

Fig. 8 Comparison between core data and calculated porosity,


water saturation, permeability, and rock-fabric values from
petrophysical analysis of SSAU 2309.
SPE Reservoir Evaluation & Engineering, April 1998

Fig. 9 Eastwest cross section of permeability distribution.

Fig. 10 Three-dimensional image of porosity distribution of


Cycles 8 and 9 in Seminole San Andres Unit, Tract 23-28.

geologic features such as upward-shoaling sequences cannot be


found; and the stochastically generated permeability data with
stratigraphic constraints largely preserves the features of upwardshoaling sequences (Fig. 11c).
Scale-Up For Flow Model

One of the objectives of the 3D reservoir modeling is to generate


flow models at various scales for reservoir simulation. To generate
flow models, it is necessary to scale-up small-scale data into larger
simulation blocks in both horizontal and vertical directions.
Scale-up in the horizontal direction is discussed by Journel and
Huijbregts10 and Perez and Kelkar11 using horizontal well data, and
by Wang et al.4 using Lawyer Canyon outcrop data. Scale-up in the
vertical direction is related to topics of vertical permeability and the
ratio of vertical permeability to horizontal permeability. Core
analyses commonly show a ratio of vertical to horizontal permeability ranging from 0.1 to 1. This general trend holds in SSAU core
data, as illustrated in Fig. 12a, by data from the SSAU 2710 well.
Open circles are data from Cycle 9, squares are data from Cycle 10,
crosses are data from Cycle 11, and solid circles are data from Cycle
12. Statistically speaking, the average k vh ratio is about 0.3.
Many formulas have been proposed for scale-up of the k vh ratio
from core data to simulation block sizes.12,13 Wang et al.4 applied
a simple analytical equation to illustrate interesting features in
scale-up of the vertical permeability of carbonates (Fig. 12b). The
k vh (k# h /k# v ) ratio decreases with the vertical gridblock size up to 20
ft (6.1 m) to a value of 0.06 and remains at a constant value at a
vertical gridblock size greater than 20 ft (.6.1 m). This limiting
SPE Reservoir Evaluation & Engineering, April 1998

Fig. 11Permeability distribution in an east-west cross sections


of (a) linearly interpolated model, (b) stochastic realization without stratigraphic constraints, and (c) stochastic realization with
stratigraphic with stratigraphic constraints.

value of 20 ft (6.1 m) is close to the average thickness of HFCs


and suggests that data variance increases significantly within a
cycle but only slightly among cycles.
Reservoir Simulation

Reservoir simulations were performed using outcrop and subsurface models to study critical factors affecting recovery efficiency.
Factors studied are geometry and distribution of rock-fabric units,
direction of water injection, the k vh ratio, dense mudstone distribution, initial gas cap, and stochastic realizations.
Lawyer Canyon Outcrop. The flow model for the Lawyer Canyon window (Fig. 13a) was constructed by overlaying the rockfabric units on the stratigraphic framework and by assigning each
unit an average porosity and permeability (Table 2).3 The result is
a geologically constrained description of the spatial distribution of
petrophysical properties. Grainstone flow units in Cycles 1 and 2
are continuous, and permeability is high throughout the entire
model, whereas in Cycle 9 the grainstone flow unit appears only in
the south-central part. A number of two-dimensional waterflood
experiments were conducted using this model to show the large
impact of the geometry and distribution of rock-fabric facies3,8 and,
particularly, the impact of low-permeability mudstone layers on
performance predictions.
Geometry and Distribution of Rock-Fabric Units. Senger et al.3
demonstrated the importance of the correct spatial permeability
distribution by comparing simulation results using the outcrop
model (Fig. 13a) with results using a simulated subsurface model
109

Fig. 12(a) Vertical permeability vs. horizontal permeability


sorted by high-frequency cycles, and (b) kvh as a function of
vertical block size.

constructed by the linear interpolation of cycles and petrophysical


data between the ends of the model (Fig. 13b). The difference in the
predicted recovery is about 13% of the OOIP, or 48% recovery from
the interpolated model and 35% from the outcrop model (Fig. 13c)
because the high-permeability grainstone in the middle of Cycle 9
does not extend to either end and was missing in the linearly
interpolated model.
The second experiment illustrated how the position of wells
relative to the spatial distribution of permeability affects recovery.3
In Fig. 14a and b, water saturation and recovery of two runs,

Fig. 13Lawyer canyon flow models. (a) The rock-fabric permeability model based on continuous outcrop data. (b) A linear
interpolation of permeability data taken from two pseudo-wells
on either end of the Lawyer Canyon window. (c) Comparison of
waterflooding performance between two models. The rockfabric permeability model gives lower recovery than the linearly
interpolated model because the permeable grainstone unit is
missing in the linearly interpolated model.

TABLE 2PROPERTIES OF ROCK-FABRIC FLOW UNITS FOR LAWYER CANYON OUTCROP MODEL
(FROM SENGER ET AL.3)
Flow
Unit

Rock Fabric

Porosity

Permeability
(md)

Initial Water
Saturation

Residual Oil
Saturation

1
2
3
4
5
6
7
8
9
10
11

Mudstone
Wackestone
Grain-dominated packstone I
Grain-dominated packstone II
Grain-dominated packstone III
Moldic grainstone I
Moldic grainstone II
Highly moldic grainstone
Grainstone I
Grainstone II
Grainstone III

0.040
0.105
0.085
0.129
0.118
0.145
0.159
0.230
0.095
0.110
0.135

0.01
0.30
4.50
1.80
5.30
0.70
2.20
2.50
9.50
21.3
44.0

0.900
0.405
0.214
0.400
0.243
0.091
0.077
0.041
0.189
0.147
0.103

0.01
0.40
0.35
0.35
0.35
0.40
0.40
0.40
0.35
0.25
0.25

110

SPE Reservoir Evaluation & Engineering, April 1998

This k vh effect on waterflooding is illustrated in Fig. 15, which


shows water saturation distributions after 20-year waterflooding for
three cases: k vh 5 0, 0.1, and 1. When k vh 5 0, the noncommunicating case, severe channeling occurs through high-permeability flow units. When k vh 5 0.1, crossflow increases and Cycles
3 to 7 are better swept than in the noncommunicating case. When
k vh 5 1, increased crossflow in Cycles 3 to 7 improves the sweep
and recovery efficiencies.
Simulation results of these cases are summarized in terms of
recovery efficiency with respect to pore volume of water injected
(Figs. 16a and b). For all k vh ratios, recovery is higher in the model
without dense mudstone layers. In Fig. 16c, recovery efficiencies
at a 0.3 pore volume of water injected in both models are plotted
with respect to the k vh ratio. The upscaled k vh ratios for the model
without dense mudstone layers, corresponding to the same recovery
in the model with dense mudstone layers at k vh ratios of 0.3 and 1,
are, respectively, 0.02 and 0.04. The k vh ratio range of 0.02 to 0.04
agrees well with the k vh ratio reported in most large-scale reservoir
simulations,4,14 where coarse-scale simulation grids are used and
dense mudstone layers are averaged in. Therefore, adding dense
mudstones to the flow model reduces the crossflow resulting from
the artifact of scaling up from fine-scale reservoir models to
coarse-scale simulation models.
All the outcrop simulations were conducted using a dead oil
without solution gas. However, the SSAU crude has a high initial
solution gas ratio and the field had a small initial gas cap. It is

Fig. 14 Water saturation distribution after 24 years of water


injection in (a) a left-to-right injection experiment and (b) a
right-to-left injection experiment, showing crossflow points at
the downflow termination of high permeability in Cycle 9 and oil
left in the middle of Cycle 7. (c) Comparison of waterflooding
performances between (a) and (b).

flooding from left to right (a) and right to left (b), are compared.
In both cases, water channels through the highly permeable grainstones in Cycles 1, 2, and 9. Channeling is more severe in Cycles
1 and 2 than in Cycle 9. Water channels through the grainstone in
Cycle 9 until it is laterally terminated, where it then flows down
across the basal mudstone of Cycle 9 into underlying Cycles 8 and
7 and leaves oil in the middle of Cycle 7 unswept. Twelve percent
more oil is trapped when water is injected right to left than when
water is injected from left to right (Fig. 14c) because the upstream
barrier of Cycle 9 is shorter and water is channeling faster in case
(b) than in case (a).
Effects of kvh Ratio and Dense Mudstones. Many simulation
studies have shown that the k vh ratio is one of the most dominant
parameters affecting recovery efficiency. Determining the k vh ratio
of flow models is one of the major issues in reservoir analysis.14
The effect of k vh ratio on recovery was tested in Lawyer Canyon
outcrop models with and without dense mudstone layers. The
model without dense mudstone layers can be considered as an
analog of the coarse-scale simulation model where dense mudstone
layers are averaged in during scale-up. In each case, simulations
were run with a k vh ratio of 0.001, 0.01, 0.1, and 1.
SPE Reservoir Evaluation & Engineering, April 1998

Fig. 15Effect of kvh value on water-saturation distributions


after 20 years of water injection. The kvh values varying from (a)
0, (b) 0.1, and (c) 1.0.
111

Fig. 17Effect of kvh ratio on waterflooding recovery using a 2D


cross section model with an initial gas cap in SSAU, Tract 23-28.

units to study their effect on recovery and production and injection


rates. Results on recovery (Fig. 18a) indicate that no significant
differences occurred from the three realizations using a correlation
length of 1,000 ft (300 m). This was partly because these realizations were constrained by rock-fabric units. Nevertheless, production and injection rates are different in these runs (Figs. 18b and
18c). These differences stem from the change in permeability
distribution and effective permeability with realization.
Conclusions

Rock fabrics are defined on the basis of grain and crystal size and
sorting, interparticle porosity, separatevug porosity, and the presence or absence of touching vugs. Petrophysical properties of
porosity, permeability, relative permeability, and capillary pressure
can be grouped according to rock fabrics. Permeability profiles can
be calculated using rock-fabric-specific transforms between interparticle porosity and permeability. Special core analysis data indicate that waterflood recovery decreases and residual oil saturation
increases with increasing separate-vug porosity. Residual oil sat-

Fig. 16 Effect of dense mudstones on the kvh value. (a) Lawyer


Canyon outcrop model with dense mudstones, (b) Lawyer Canyon outcrop model without dense mudstones, and (c) the technique to determine an effective kvh value for the model without
dense mudstones based on the kvh of core data.

therefore important to understand how the k vh ratio affects recovery


efficiency when SSAU crude is used and an initial gas cap is
present. This k vh effect was studied using a SSAU cross-section
model through SSAU 2316, 2602, 2506, 2502, 2505, and 2704
wells. Three simulation runs were performed using k vh 5 0, 0.1,
and 1. Recovery curves (Fig. 17) show a trend of decreasing
recovery with an increase of k vh ratio, which is opposite to the trend
observed in the outcrop models (Fig. 16). The opposite trend stems
from an initial gas cap being presented in the subsurface model. The
high-mobility gas cap serves as a big conduit for water channeling,
and increasing the k vh ratio increases the crossflow of water through
an unperforated gas cap. The k vh ratio can be estimated by matching
simulation results with production history. Wang et al.4 used a k vh
ratio of 0.04 in an 80-acre 3D model to match the SSAU production.
This suggests that the barrier effect in SSAU is strong.
Effect of Stochastic Realization. Because uncertainties in flow
models generated from well data are high, the chances of matching
field production using a linearly interpolated model are low, and
history matching may be better achieved by using stochastic models. Flow simulations were performed on stochastic realizations
conditioned on SSAU well data and constrained by rock-fabric flow
112

Fig. 18 Effect of stochastic realization on (a) waterflooding


recovery, (b) production rate, and (c) injectivity.
SPE Reservoir Evaluation & Engineering, April 1998

urations determined by the steady-state method are significantly


lower than those determined by the unsteady-state method, because
the real residual oil saturation is not reached in the fast unsteadystate method.
HFCs and rock-fabric units are the two critical scales for
modeling shallow-water carbonate ramp reservoirs. Descriptions of
rock-fabric facies stacked within HFCs provide the most accurate
framework for constructing geologic and reservoir models because
discrete petrophysical functions can be fit to rock-fabric units and
fluid flow can be approximated scale-up within rock-fabric flow units.
Stochastic simulations without stratigraphic control can be too
random for generating geologically realistic models. The upwardshallowing sequences of carbonates can be observed in stochastic
models only when they are constrained by rock-fabric flow units.
Scale-up of permeability in the vertical direction was investigated in terms of the k vh ratio. Because of the cyclic nature of
carbonate reservoirs, the k vh ratio decreases exponentially with the
vertical gridblock size up to the average cycle size of 20 ft (6.1 m)
and remains at a value of 0.06 for a gridblock size of more than 20
ft (.6.1 m).
Simulation results showed that critical factors affecting recovery
efficiency are the stacking patterns of rock-fabric units, the k vh
ratio, and the dense mudstone distribution. Simulations using
outcrop models demonstrated that the geometry and distribution of
rock-fabric units can significantly affect recovery efficiency and the
direction of water injection. The k vh ratio and dense mudstone
layers are the primary controlling parameters governing the recovery efficiency. Waterflood recovery increases with the k vh ratio
when dead oil is used and decreases with the k vh ratio when an
initial gas cap is present. When stochastic models are constrained
by the rock-fabric framework, simulation results are similar in
recovery but different in production and injection rates.
Nomenclature
h 5 thickness, m
k 5 permeability, md
k oi 5 permeability at initial oil saturation, md
k ro 5 relative oil permeability, fraction
k rw 5 relative water permeability, fraction
k h 5 horizontal permeability, md
k# h 5 arithmetic mean of horizontal permeability, md
k v 5 vertical permeability, md
k# v 5 harmonic mean of horizontal permeability, md
k vh 5 ratio of vertical permeability to horizontal permeability, fraction
R vp 5 ratio of separate-vug porosity to total porosity, fractions
Acknowledgments

This research was done at the Reservoir Characterization Research


Laboratory of the Bureau of Economic Geology and was funded by
industrial sponsors and by DOE Contract no. AC22-89BC1440.
Publication was authorized by the Director, Bureau of Economic
Geology, The U. of Texas at Austin. We thank Susan Lloyd for
word processing and layout, and Tari Weaver and David M.
Stephens, under the direction of Joel L. Lardon and Richard L.
Dillon, for preparation of illustrations.
References
1. Kerans, C. et al.: Characterization of Facies and Permeability Patterns
in Carbonate Reservoir Based on Outcrop Analogs, The U. of Texas
at Austin, Bureau of Economic Geology, final report, Contract No.
DE-AC2289BC14470 for the Assistant Secretary for Fossil Energy,
U.S. DOE, Washington, DC (1993) 160.
2. Galloway, W.E. et al.: Atlas of Major Texas Oil Reservoirs, The U. of
Texas at Austin, Bureau of Economic Geology (1983) 139.
3. Senger, R.K. et al.: Dominant Control on Reservoir-Flow Behavior in
Carbonate Reservoirs as Determined from Outcrop Studies, Reservoir
Characterization III, B. Linville et al. (eds.), Proc., Third International
Reservoir Characterization Technical Conference, Tulsa (November 1993).
4. Wang, F.P. et al.: Critical Scales, Upscaling, and Modeling of ShallowWater Carbonate Reservoirs, paper SPE 27715 presented at the 1994
SPE Reservoir Evaluation & Engineering, April 1998

SPE Permian Basin Oil and Gas Recovery Conference, Midland, Texas,
1618 March.
5. Lucia, F.J. et al.: Fluid-flow Characterization of Dolomitized Carbonate-Ramp Reservoirs: San Andres Formation (Permian) of Seminole
Field and Algerita Escarpment, Permian Basin, Texas and New Mexico, Hydrocarbon Reservoir Characterization: Geologic Framework
and Flow Unit Modeling, E.L. Stoudt and P.M. Harris (eds.), SEPM
Short Course No. 34 (1995) 129153.
6. Lucia, F.J.: Petrophysical Parameters Estimated from Visual Descriptions of Carbonate Rocks: A Field Classification of Carbonate Pore
Space, JPT (March 1983) 629; Trans., AIME, 275.
7. Lucia, F.J.: Rock-Fabric/Petrophysical Classification of Carbonate
Pore Space for Reservoir Characterization, American Assn. of Petroleum Geologists Bull. (1995) 79, No. 9, 1275.
8. Kerans, et al.: Integrated Characterization of Carbonate Ramp Reservoirs Using Permian San Andres Formation Outcrop Analogs, American Assn. of Petroleum Geologists Bull. (1994) 78, No. 2, 181.
9. Lucia, F.J. and Conti, R.D.: Rock Fabric, Permeability, and Log
Relationships in an Upward-Shoaling, Vuggy Carbonate Sequence, The
U. of Texas at Austin, Bureau of Economic Geology, Geological
Circular 87-5 (1987) 22.
10. Journel, A.G. and Huijbregts, Ch.J.: Mining Geostatistics, second edition, Academic Press, San Diego (1978) 599.
11. Perez, G., and Kelkar, M.: Assessing Distributions of Reservoir Properties Using Horizontal Well Data, Reservoir Characterization III, B.
Linville et al. (eds.), Proc., Third International Reservoir Characterization Technical Conference, Tulsa (November 1991).
12. Lishman, J.R.: Core Permeability Anisotropy, J. Cdn. Pet. Tech.
(AprilJune 1970) 79.
13. Haldorsen, H.H. and Lake, L.W.: A New Approach to Shale Management in Field-Scale Models, SPEJ (1984) 447; Trans., AIME, 277.
14. Harpole, K.J.: Improved Reservoir CharacterizationA Key to Future
Reservoir Management for the West Seminole San Andres Unit, JPT
(November 1980) 2009; Trans., AIME, 269.

SI Metric Conversion Factors

acre 3 4.046 873


ft 3 3.048*
psi 3 6.894 757
sq mile 3 2.589 988

E103
E201
E100
E100

5 m2
5m
5 kPa
5 km2
SPEREE

*Conversion factor is exact.

Fred P. Wang is a senior reservoir engineer at PGS Reservoir, Inc.,


in Houston and previously worked at the Bureau of Economic
Geology, The U. of Texas at Austin. He holds an MS degree from
The U. of Texas at Austin and a PhD degree from Stanford U.,
both in petroleum engineering. F. Jerry Lucia is a Senior Research Fellow with the Bureau of Economic Geology developing new techniques and methods for characterizing carbonate
reservoirs to improve recovery from existing oil fields through the
integration of geological, petrophysical, engineering, and production data. Previously Lucia was a Consulting Geological
Engineer for Shell Oil Co. assigned to the Head Office staff when
he retired in 1985 with 31 years experience as a geological
engineer in research and operations. Currently, he holds a BS
degree in engineering and an MS degree in geology from the
U. of Minnesota. Charles Kerans is a senior research scientist in the
Bureau of Economic Geology. He holds a BS degree from St.
Lawrence U. and a PhD degree from Carleton U., both in geology.

Wang

Lucia

Kerans
113

You might also like