You are on page 1of 10

IPTC 11199

Reactive Transport Models of Limestone-Dolomite Transitions: Implications for


Reservoir Connectivity
Yitian Xiao, ExxonMobil Upstream Research Company, and Gareth D. Jones,* ExxonMobil Development Company

*Present Address: Chevron Energy Technology Company, San Ramon, California, U.S.A

Copyright 2007, International Petroleum Technology Conference


spatial distribution of petrophysical properties and viable
realizations of reservoir connectivity for flow simulation.
This paper was prepared for presentation at the International Petroleum Technology
Conference held in Dubai, U.A.E., 4–6 December 2007.

This paper was selected for presentation by an IPTC Programme Committee following review
INTRODUCTION
of information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the International Petroleum Technology Conference
and are subject to correction by the author(s). The material, as presented, does not necessarily Diagenetic transitions between limestone and dolomite are a
reflect any position of the International Petroleum Technology Conference, its officers, or
members. Papers presented at IPTC are subject to publication review by Sponsor Society
critical control on reservoir quality in many carbonate
Committees of IPTC. Electronic reproduction, distribution, or storage of any part of this paper reservoirs. The Jurassic Ghawar field in Saudi Arabia and the
for commercial purposes without the written consent of the International Petroleum Technology
Conference is prohibited. Permission to reproduce in print is restricted to an abstract of not Permian-Triassic North Dome field in Qatar and Iran are two
more than 300 words; illustrations may not be copied. The abstract must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, IPTC, P.O.
notable prolific Middle East examples. Dolomite can behave
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435. as a reservoir or as a baffle / barrier to flow depending on
original depositional texture, style of dolomitization and
ABSTRACT presence of anhydrite cement [1-3]. High permeability, or so-
called “Super K” dolomites with multi-darcy permeability that
Substantial volumes of world hydrocarbon resources occur in sporadically occur in thin zones (typically <5 ft) are vital
interlayered limestone-dolomite reservoirs. Diagenetic contributors to individual well and overall reservoir
variations in lithology and primary depositional texture control performance [2]. Laterally continuous low permeability ‘tight”
the magnitude and spatial distribution of petrophysical dolomites can also have significant impact on subsurface flow
properties. The frequency and nature of limestone-dolomite and thus reservoir management strategies. The transition
transitions that define flow units and baffles/barriers are between limestone and dolomite is readily identified in both
critical for understanding reservoir connectivity and wireline logs and core and is dominantly observed as a sharp
optimizing field development. interface described as a “dolomite front” sensu Wilson et al.
[4]. Unfortunately, dolomite fronts in carbonate reservoirs are
Existing subsurface predictions are largely based on often below seismic detection, particularly those associated
observation and are occasionally linked to sequence with ‘thin’ beds of dolomite on the order of a few to several
stratigraphy. This approach can be relatively successful at tens of feet thick. Predicting the spatial distribution or
predicting general trends in limestone and dolomite connectivity of dolomite and associated petrophysical
occurrence, but there is considerable uncertainty in predicting properties between wells and away from well control is thus a
and correlating spatial variations in diagenetic styles at the fundamental challenge in carbonate reservoir characterization.
field scale. In contrast to depositional facies there are no defined rules or
guidelines for correlating dolomite bodies in the subsurface.
Reactive transport models that explicitly couple fluid flow and Existing correlations of dolomite tend to be lithostratigraphic
chemical reactions to facilitate quantitative investigations of and contain considerable uncertainty in the degree of
limestone-dolomite transitions are a recent addition to the interpreted reservoir connectivity (Figure 1).
predictive diagenesis tool kit. The approach is illustrated with
a generic study designed to investigate the fundamentals of Understanding the distribution and connectivity of
brine reflux dolomitization. Results provide new insights on: sedimentary bodies has been greatly enhanced from studies of
(1) Dynamic propagation of limestone-dolomite fronts by outcrop analogs [5-6]. A vast number of outcrop studies have
fingering, (2) The connectivity of dolomite ‘fingers’ to the focused on dolomitized carbonates. However, relatively few
‘main body’; (3) Dolomitization and anhydrite occurrence have specifically focused on transitions between limestone and
(primary precipitation and cements); (4) Locating reservoir dolomite and / or described the connectivity of dolomite
‘sweet spots’ in a dolomite geobody and (5) The interaction bodies. Outcrop based studies by Wilson et al [4] and Yao and
and importance of semi-regional versus local brine reflux Demicco [7] that mapped dolomite fronts associated with
flow. This process-based approach, in combination with geothermal circulation and ascending hydrothermal fluids,
available observational data and sequence stratigraphic respectively, are notable exceptions. These analogs are,
paleoenvironmental reconstructions, has significant however, not appropriate for reservoirs where the majority of
implications for predicting limestone-dolomite transitions, the dolomite is interpreted to have formed by the reflux of brines
2 !PTC 11199

from evaporated seawater, for example Ghawar and the North (A) Limestone (B)
Dolomite
Dome fields in the Middle East [2-3]. Dolomite fronts mapped Dolomite

in Plio - Pleistocene outcrops on Bonaire by Lucia and Major


[8] are interpreted to have formed by brine reflux (Figure 2).
The Bonaire study provides insights on the propagation of a Limestone
dolomite front and the distribution of petrophysical properties
in a reflux system but is of limited use in helping develop (C) (D)

dolomite correlation strategies in extensive carbonate ramp Limestone

reservoirs like Ghawar and the North Dome. Other smaller


scale outcrop examples of limestone dolomite transitions
interpreted to have formed as a consequence of brine reflux Dolomite
are shown in Figure 3.

Type 1: Dolomite concordant with stratigraphy


Figure 3. Outcrop examples of limestone dolomite transitions. (A)
Devonian, Cripple Creek, Canada; (B) Ordovician, Franklin
Mountains, USA, after Tillotson [9]; (C) and (D) Ordovician,
Franklin Mountains, USA (Images courtesy of Jemery Bellian).
In all images Dolomite is tan (light brown) and limestone is grey.
Red arrow shows interpreted direction of brine reflux paleo fluid
flow.

Reactive transport models explicitly couple flow and chemical


reactions facilitate investigations of the distribution of
diagenetic alteration, including dolomitization, in the
Type 2: Dolomite discordant with stratigraphy subsurface [10-12]. Numerical models have the advantage of
being able to simulate the specific system of interest and to
track the spatial and temporal evolution of diagenesis in
contrast to the terminal diagenetic overprint observed in
outcrop studies [13-15].

In the past three years, we have applied reactive transport


models to investigate several styles of diagenesis in both
0 10 km carbonate and siliciclastic rocks: A) Early diagenesis in a
freshwater lens and associated hydrological zones in an
isolated carbonate platform; B) Geothermal convection and
Figure 1. Correlations of limestone to dolomite (shaded grey) burial diagenesis in a salt buried isolated platform; C) Fault
transitions from wireline well logs in the Ghawar Field, Saudi induced hydrothermal fluid flow and illitization; and D)
Arabia. Note 1-2 km well spacing and vertical scale in ft. After Formation damage associated with steam (water) injection and
Cantrell et al. [2]. artificial diagenesis (Figure 4). Adopting a sensitivity analysis
approach, we examined how key natural variables, such as
climate, sea level, temperature, salinity and fluid composition
and porosity/permeability heterogeneity impact different styles
depositional dip of diagenesis and reservoir evolution. We also examined how
formation damage caused by artificial diagenesis in production
can be predicted by reactive transport modeling. Linking
fundamental geochemical processes to predict reservoir
quality has significantly improved our understanding of
several conceptual diagenetic models and paradigms. Our
novel approach suggests that reactive transport models, when
sufficiently integrated with traditional methods and calibrated
with field data, have the potential to significantly improve our
0 10 km
ability to predict carbonate and siliciclastic reservoir quality
[11-12, 15-16].

Figure 2. Limestone to dolomite transition mapped in a Plio In this paper we present results from reactive transport models
Pleistocene outcrop on Bonaire, Netherlands, Lesser Antilles. employed to investigate brine reflux dolomitization with an
Note dolomite ‘fingering’ in direction of depositional dip. emphasis on investigating the origin, distribution and
Modified after Lucia and Major [8]. connectivity of dolomite bodies. We present a brief synthesis
of our previously published 2D reactive transport model study
IPTC 11199 3

of reflux dolomitization [11] followed by preliminary results For example, the reaction of dolomitization can be expressed
from 3D simulations that provide new insights on the as:
distribution of dolomite, anhydrite and implications for
reservoir connectivity and subsurface correlation strategies. ∂Cdolo − a
E
Q
= S A e RT ( − 1) 2.26 (2)
∂t K eq
(A) (B)
25
W 25
30
35 30
k h f i Where S represents the reactive surface area, A is the rate
constant, Ea is the activation energy, Q/Keq is the saturation
40
35
45 40

index, and 2.26 is the reaction order [17].


50 45

Porosity 55 50
45
60 60
65 55 50
70
Vadose
S.L. .2 km
calcite cement, % 50 Due to complex boundary conditions and complicated
Karst 2 km –2 0 2
5
cm/yr coupling between the transport and reaction terms, it is
(C) (D)
Calcite Cement
impossible to provide analytical solutions to equation (1) for
even the simplest geochemical system. Therefore numerical
Illite solutions have to be used. Fortunately, due to the exponential
increase in computational power, realistic reactive transport
models have been developed that are beginning to provide
exploitable insights to reservoir quality prediction. We have
used two commercially available numerical programs, Xt2
Illite kaolinite Kaolinite [18] for the 2D models and ToughReact [19] for the 3D
models. The rest of this paper highlights results from our
recent efforts in applying this emerging technology to the
quantitative prediction of the spatial and temporal distribution
Figure 4. Previous reactive transport modeling applications: (A)
Early diagenesis in an isolated carbonate platform; (B)
of dolomitization and its effect on carbonate reservoir quality
Geothermal convection and burial diagenesis in a salt buried and connectivity.
isolated platform; (C) Fault induced hydrothermal fluid flow and
illitization; and (D) Formation damage associated with water RESULTS
injection. After Xiao and Jones [12].
2D Simulations of Brine Reflux Dolomitization
NUMERICAL MODELS AND APPROACH
Our 2D model was based on the reflux model originally
A reactive transport model relies on a mathematical proposed by Adams and Rhodes [20] to explain dolomitization
formulation to describe geochemical processes involving of Permian carbonate in West Texas, which shows a close
fluid-rock interactions. The general governing equation can be stratigraphic association with platform interior evaporites.
written as: The reflux model has been modified by several researchers
[21-24] to incorporate different geological scenarios recently.
∂ ∂ ⎛ ∂C ⎞ ∂C ⎛ ∂C ⎞ (1) Our 2D simulation domain is 10 km x 500 m (Figure 6). The
(φC i ) = ⎜ φD i ⎟ − φv i + φ ∑ ⎜ i ⎟
∂t ∂x ⎝ ∂x ⎠ ∂x k ⎝ ∂t ⎠ k
simulation domain is divided by 10 x 5 nodes in x- and y-
directions. The upper and lower boundaries are specified as no
Where Ci is the concentration of a specific species in the pore flow (a constraint of current version of the Xt2 program) and
fluid, D is the combined diffusion and dispersion coefficient the left (platform interior) and right (basin ward) boundaries
term, v is the linear fluid flow rate, and ф is the porosity. The are open to flow. In our 3D simulation (discussed in the next
first two terms on the right hand side describe the transport section) the upper boundary is represented as a fluid pressure
process (diffusion, dispersion, and advection) while the last boundary allowing both the recharge and discharge of brines
term describes the effect of geochemical reactions (Figure 5). on a submerged platform-top [25]. Our simulated flow domain
thus represents the section of a carbonate platform where brine
recharge and predominantly horizontal reflux flow occurs as
demonstrated in the Grosmont platform [26]. Reflux flow,
Simulation tracks fluid flow which is from the left (platform interior) to the right (basin
and chemical reactions in wards) was set by specifying injection of brine at a rate of 5
Inflow porous media
Outflow
m3/s in the uppermost left cell (interior) of the flow domain.
The initial flow velocity in the majority of the flow domain is
modify porosity/permeability
Mg2+ Ca2+ 3-4 m/yr (at shallow depth) but increases to ~6 m/yr close to
the brine source. The initial flow velocity is slower (< 3 m/yr)
Initial rock property
in a zone that begins 150 m below the brine source and
Dolomitization: 2CaCO3 + Mg2+ = CaMg(CO3)2 + Ca2+ extends 5.5 km basin ward (Figure 6). The flow velocities in
this initial flow field are relatively high but consistent with
Figure 5. Schematic diagram showing the reactive transport previous simulations of reflux in carbonates in the shallow (<
process involving dolomitization. 500 m) subsurface [26].
4 !PTC 11199

Brine injection No Flow (A) Initial Porosity


node
Open

Open
(B) Initial Permeability

Figure 6. Simulation flow domain, hydraulic boundary


conditions, simulation grid and location of brine injection node. (C) Initial Fluid Velocity
Initial distribution and magnitude of reflux fluid velocities
(contours are of fluid velocity in m/yr with a contour interval of
0.5). After Jones and Xiao [11].

To capture the heterogeneous porosity and permeability of


carbonate sediments in the shallow subsurface (0-500 m), due
to either variations in sediment texture or early diagenesis [27-
28], we incorporated heterogeneous porosity and permeability Figure 7. Initial porosity, permeability and resultant flow field
distributions in our 2D models to investigate the effect of used for investigating the effect of porosity and permeability
heterogeneity on reflux dolomitization. The porosity is treated heterogeneity (see Figure 7). (A) Porosity (contours are porosity
as a field variable in the simulation. We set it to a random [fraction] with a contour interval of 0.01; (B) Permeability
distribution about 35%, with a standard deviation of 2.5% (i.e., contours are of horizontal permeability [Darcys] with a contour
porosity is 0.35 ± 0.25). We placed a top threshold of 35 % on interval of 20); (C) fluid velocity (contours are of fluid velocity
this porosity distribution to avoid uncharacteristically high [m/yr] with a contour interval of 0.5). After Jones and Xiao [11].
values of permeability that would promote numerical
instability. Sampling randomly from this truncated porosity The simulation results demonstrated that the initial porosity
distribution generated a heterogeneous porosity distribution and permeability heterogeneity in the precursor sediment had a
that ranged from 29 to 35 % (Figure 7A). We specified the significant effect on the pattern of dolomitization (Figure 8).
Class 1 relationship [11] between porosity and permeability to The applied heterogeneity results in the formation of
generate a corresponding heterogeneous permeability pronounced perturbations, in the geometry of the dolomite
distribution, which ranged from 12.8 to 126 Darcies (Figure front, which extended several km basin ward from the main
7B). The permeability anisotropy was held at 100. dolomite body at 2 My (Figure 8 A-E). These perturbations,
described as dolomite “fingers” by Wilson et al. [4], became
To enhance the spatial heterogeneity and investigate smaller more pronounced with time suggesting a positive feedback
scale features of the reflux process, we doubled the number of between the initial exploitation of preferential flow paths and
nodes in the flow domain by decreasing the node spacing to subsequent changes in rock properties by dolomitization that
500m in the horizontal direction and 50 m in the vertical enhance flow (Figure 8 A-E) as well as negative feedback
direction (20x10 nodes in x- and y-directions). We reduced the associated with anhydrite cementation (Figure 8 F-J).
initial rate of reflux flow to a maximum of 0.9 m/yr (range 0.1
to 0.9 m/yr) to offset the increase in computation time (by a The simulation results indicate the association of anhydrite
factor of 8) associated with an increase in the number of grid cementation and dolomite distribution was partially and
nodes (Figure 7C). Our simulated range of porosity and in temporally (Figure 8 F-J). The initial anhydrite cement was
particular permeability heterogeneity, which can range up to formed due to the interaction of SO42- from the brine with the
five orders of magnitude in shallow carbonate sediments [27- excessive Ca2+ released from the limestone due to
28], is relatively moderate. No explicit coupling of multiple dolomitization. However, as the dolomitization completed
deposition and diagenesis stratigraphic sequences “upstream”, no more excessive Ca2+ was released, which led
(parasequence stacking) was incorporated in our simulated the brine dipping below gypsum (anhydrite) saturation. This is
heterogeneity. The total simulation time is 2 m.y. the primary reason for the remobilization of the initially
formed anhydrite cementation (Figure 8 F-J). Focusing of
IPTC 11199 5

reflux flow resulted in enhanced anhydrite dissolution in the distribution of petrophysical properties. 2D reactive transport
dolomite fingers, whereas anhydrite had a greater preservation models of dolomitization are limited by their ability to
potential adjacent to the fingers (Figure 8 J). Dolomitization incorporate spatial variations in natural variables that control
enhances porosity by up to 8 % and anhydrite cements occlude brine reflux and thus extrapolation to the distribution of
porosity by up to 14 % (Figure 8 K-O). The dolomite fingers dolomite in 3D. We employed 3D reactive transport models to
were porosity “sweet zones” surrounded by non- to partially investigate the effect of permeability and spatial distributions
dolomitized sediments that were locally cemented with in the magnitude and distribution of platform-top brines that
anhydrite and have a lower porosity (Figure 8 K-O). source subsurface brine reflux. Salinity variations observed in
modern environments, for example Lake Macleod [29] in
Dolomite Anhydrite Porosity addition to those interpreted from the thickness and type of
evaporite deposits in the rock record were used to condition
(A) 0.2 My (F) 0.2 My (K) 0.2 My

model boundary conditions and initial rock and fluid


properties. We specified a flow domain 5 km x 5km x 100 m
with a grid node spacing of 200 m horizontally and 2 m
(B) 0.32 My (G) 0.32 My (L) 0.32 My

vertically. Reflux was simulated from three platform-top brine


ponds approximately 500 m in diameter with different
(C) 0.56 My (H) 0.56 My (M) 0.56 My
maximum salinities of 210, 175, and 140 ppt (Figure 9) The
basal boundary was specified as a no flow boundary (refluxing
brines never reached this depth in the geological time
(D) 1 My (I) 1 My (N) 1 My simulated) and the sides of our flow domain cube were
specified as hydrostatic and thus open to flow. All other
parameters and boundary conditions were specified as those
(E) 2 My (J) 2 My (O) 2 My used in our 2D reactive transport models of brine reflux that
incorporated a heterogeneous distribution of permeability
described above in this paper and [11].

Model results showing the 3D evolution of brine reflux up to


5000 years are depicted in Figure 9. Individual reflux brine
Figure 8. Effect of initial porosity and permeability heterogeneity plumes sourced by the three different ponds, evident after 500
(see Figure 7) on dolomite, anhydrite and porosity (A-E) yrs of flow, eventually coalesce to form a single brine plume
Dolomite (contours are percent dolomite with a contour interval (Figure 9). At 5000 years the mesosaline to hypersaline brine
of 5); (F-J) anhydrite (contours are percent anhydrite with a plume extends over the 25 km2 area modeled to a maximum
contour interval of 2); and (K-O) porosity (contours are porosity depth of approximately 50 m but with considerable variation
[fraction] with a contour interval of 0.02). After Jones and Xiao
[11].
in vertical extent related to brine pond proximity and local
permeability heterogeneity. Modeled maximum rates of brine
Our 2D reactive transport simulations demonstrate the reflux were on the order of few meters per year with spatial
complex and dynamic nature of dolomitization associated variations controlled by the proximity to the different brine
anhydrite cementation and porosity evolution in a simplified ponds and permeability. The resultant flow field is the net
reflux system. Incorporation of initial porosity and product of both reinforcing and opposing flows from the three
permeability heterogeneity resulted in a complex dolomite separate ponds. In contrast to uni-directional flow in our 2D
geometry with pronounced dolomite “fingers” extending results, 3D simulations demonstrate that refluxing bines, at
ahead of the main dolomite body. The simulated dolomite least locally in the vicinity of brine ponds, have the potential
distributions seem to change from thick and extensive to thin to flow up and down depositional dip and along depositional
and separated in the direction of the reflux flow, consistent strike. After approximately 5000 yrs the brine plume appeared
with outcrop observations [Figures 2 and 3]. The predicted to reach a quasi steady state relative to the effective hydraulic
distribution of limestone to dolomitization transition and head in the brine ponds that drives flow. Beyond 5000 years
potential location of reservoir “sweet” spots can have advection is limited and the slower process of diffusion
important implications on reservoir connectivity study. For controls brine plume spreading. It is unlikely that steady state
example, the extensive main dolomite body in the platform reflux would be obtained in nature due to variations in climate,
interior (Figure 8 E) suggests good lateral and vertical sea level and geomorphology that act at shorter time scales to
reservoir connectivity, whereas in the limestone to dolomite change the distribution of platform-top salinity that drives
transition zone the layered dolomite “fingers” and anhydrite flow.
cement indicates complex reservoir connectivity, with good
reservoir intervals (uncemented dolomite) sandwiched by tight The pattern of brine reflux observed in our 3D simulations
zone with extensive anhydrite cemented (Figure 8 J, O). generated more complex distributions of dolomite and
anhydrite than observed in our 2D study and has significant
3D Simulations of Brine Reflux Dolomitization implications for understanding connectivity in carbonate
Geological models developed for simulating production of reservoirs (Figures 10 and 11).
hydrocarbons and reservoir management requires 3D
6 !PTC 11199

50 yr 100 yr Thus reflux associated with local brine ponds has the potential
• 5 km x 5 km x 100 m
to generate complex distributions of dolomite with ‘finger’
• Three brine ponds orientations that reflect multi directional fluid flow. This has
• Regional brine flow
three important implications for subsurface correlation and
geological modeling of reflux dolomites: 1) Understanding the
• Salinity 4~5 x seawater processes of reflux and paleo fluid flow can provide a
framework for correlation scenarios and guidelines for
Salinity 5000 yr
200 yr geological modelers 2) Reflux flow can be locally multi
distribution directional with respect to depositional dip and strike and 3)
Dolomite “fingers” tend to connect to the “main body” in a
single reflux episode and are rarely isolated “pods”. The
occurrence of isolated dolomite pods is restricted to the early
termination of reflux as illustrated by the distribution of
dolomite at times less than 60 k.y. (Figure 10), multiple
500 yr
episodes of brine reflux separated in time and space and
dolomitization by other mechanisms [12].

(A) (B) Dolomite


main body

20 ky 40 ky
Figure 9. 3D Simulation flow domain and boundary conditions
and distribution of brine ponds and RTM 3D distribution of Dolomite 60 ky
salinity and evolution of reflux brines at 50, 100, 200, 500, and geobody Dolomite
fingers
5000 yr. Sea water in black, hypersalinity brines in yellow and
orange, and mixed fluids in blue.
100 ky
80 ky
Dolomite
main body
3D Evolution of Dolomite - ‘Main Body’ and ‘Fingers’ (D) (C)

The evolution of dolomite, simulated over a period of 100 ky


for the reflux brine plume depicted in Figure 9, demonstrates Dolomite
how initially small isolated dolomite bodies located beneath fingers

and adjacent to the brine ponds grow and amalgamate to form


a single dolomite body that extends over the entire 25 km2
(Figure 10). Results show dolomite bodies grow at different
rates. In a similar manner to our 2D simulations the rate of Figure 10. RTM 3D distribution of dolomite: (A) Dolomite
evolution from 20 to 100 k.y.; (B)-(C) 2D slices showing the
dolomite body propagation is a function of the Mg2+
dolomite main body and fingers in x- and y-directions; (D) 2D
concentration of the brine and the flow rate. In a reflux system slices showing dolomite in z-direction. Dolomite in red and
this is controlled by the salinity of the brine ponds, distance limestone in blue.
from the pond and the magnitude of permeability (assuming a
uniform temperature and reactive surface area). Simulated 3D Evolution of Anhydrite Distribution (Bedded and
dolomite body growth occurs by the propagation and eventual Cements).
amalgamation of dolomite ‘fingers’ sense Wilson [4]. The
reaction front from limestone to dolomite is sharp. Dolomites generated in a reflux system have a multi faceted
Visualization of the 3D dolomite body after 100 kyr of reflux relationship with evaporites that are depositional and / or
shows the significant variation in the dimensions and shape of occur as cements [11]. Our 3D simulations of brine reflux
dolomite “fingers” that extend form the “main” dolomite dolomitization show two types of anhydrite: 1) Anhydrite
body, here described as dolomite that is laterally continuous cements that precipitate ahead of a migrating dolomite front
over the areal extent of the flow domain (Figure 10). due to the release of calcium as explained in our 2D study [11]
Dolomite fingers thin in the direction of flow and tend to and 2) the direct precipitation of “bedded” anhydrite from
become more pronounced with time depending on the evaporated seawater (Figure 11). Bedded anhydrite is
specified feedback of both dolomization and anhydrite distributed as a laterally continuous body (Figure 11) with rare
cementation on permeability and the initial permeability of isolated “holes” that may act as potential fluid conduits
limestone the dolomitization front intersects along a reflux (Figure 11 D). Anhydrite cement has a more complex
flow path. Examination of model results using 2D slices show distribution that is related to the distribution of dolomite.
dolomite bodies that are apparently isolated from the main Greater preservation of anhydrite cement is evident in the 3D
body (Figure 10). However, these bodies are connected to the models of dolomitization (compare Figures 8 and 11). This is
main body in 3D and were generated by the transport of Mg2+ due to the shorter duration of brine reflux simulated and the
along a flow path out of the 2D plane visualized (Figure 10). precipitation of bedded anhydrites that reduce permeability,
IPTC 11199 7

retard flow and slow rates of anhydrite dissolution. Future


simulations will explore the relationship between dolomization (A)
Dolomite
(B) Anhydrite
main body
and anhydrite and the implications for identifying dolomite bedding

reservoir sweet spots where anhydrite cements are rare /


absent.

(A) (B) Isolate


Anhydrite Anhydrite
dolomite cement
bedding
fingers
20 ky 40 ky

60 ky
Anhydrite
geobody
(C)

Anhydrite
cement
100 ky
80 ky
Anhydrite High
bedding Permeability
(D) (C) Dolomite

Anhydrite
cement

Figure 12. RTM 3D distribution of (A) dolomite; (B) Anhydrite;


(C) High-permeability dolomite (dolomite with < 5% anhydrite
Figure 11. RTM 3D distribution of anhydrite (bedded and and > 15 % porosity). All properties are shown in 2D slices in x-
cement): (A) Anhydrite evolution from 20 to 100 k.y.; (B)-(C) 2D direction. Porosity intervals are shown in contours and “high-K”
slices showing the bedded anhydrite (red) and anhydrite segment dolomite is shaded in red. Thin section photo courtesy of Jim
(green) in x- and y-directions; (D) 2D slices showing anhydrite in Markello.
z-direction. Thin section photos courtesy of Jim Markello.
DISCUSSION
High Permeability Dolomite Distribution
Reactive transport models have improved our fundamental
Understanding the distribution and connectivity of dolomite understanding of reflux dolomitization. Results provide a
bodies in the subsurface is the first step towards building more process-based framework for understanding the evolution of
accurate reservoir models. The magnitude and distribution of dolomite, associated anhydrite cements and petrophysical
permeability in a dolomite body is an equal if not more, properties. Insights from these generic simulations can be
important consideration. Contrasts in dolomite permeability applied to reservoir examples to help constrain and develop
can be extreme and occur over short distances. The problem is scenarios for subsurface correlations of dolomite bodies and
exemplified by the occurrence of high permeability (> 100 mD their connectivity.
to several D) dolomites that are often thin (a few feet) such as
those in Ghawar which account for a significant portion of Our 2D and 3D modeling results provide unique insight about
total production [2]. Reactive transport models provide a the dynamic propagation of limestone-dolomite fronts by
process based method of investigating the origin and spatial fingering, the connectivity of dolomite ‘fingers’ to the ‘main
distribution of dolomite permeability (Figure 8) [11]. We body’, dolomitization and anhydrite occurrence (primary
interrogated our 3D model to illustrate reservoir sweet spot precipitation and cements), and locating reservoir ‘sweet
identification applying the following illustrative criteria: spots’ in a dolomite geobody. The predicted pattern of
Dolomite with > 15 % porosity and < 5% anhydrite (Figure limestone to dolomite transition is consistent with outcrop
12). Dolomite “sweet spots” (shown in red) are located close observations (Figures 2 and 3), yet the modeling results
to the dolomite front, have a variable thickness and are offered theoretical basis for new interpretations.
generally well connected in this case (Figure 12). These sweet
spots occur in both the dolomite “main body” and “fingers” For example, the 2D simulated dolomite distributions (Figure
(Figure 12). This general pattern is similar, but not exclusive 8) seem to change from thick and extensive to thin and
to several dolomite reservoirs, although our model is a separated in the direction of the reflux flow. The predicted
simplified hypothetical example (authors’ personal distribution of limestone to dolomitization transition and
observation). Permeability reduction via the potential location of reservoir “sweet” spot revealed some
overdolomitization process was not observed in our 3D model, unique features. In particular, the extensive dolomite main
perhaps due to the relatively short simulation time. Chemical body in the platform interior (Figure 8 E) suggests good lateral
compaction, hydrothermal dolomitization and dolomite and vertical reservoir connectivity, whereas in the limestone to
recrystallization could cause overdolomitization. These dolomite transition zone the layered dolomite “fingers” and
diagenetic processes were not included in our simulations. anhydrite cement indicates complex reservoir connectivity,
with good reservoir intervals (uncemented dolomite)
8 !PTC 11199

sandwiched by tight zone with extensive anhydrite cemented The models predicted the spatial and temporal distribution of
(Figure 8 J, O). key geological parameters, such as limestone, dolomite,
anhydrite and porosity, and how these parameters were
Our 3D modeling results suggest that the extensive dolomite affected by depositional and diagenetic variables (initial
main bodies and the underline dolomite fingers (which are lithology, porosity and permeability, diagenetic fluid
often a few feet thick) are not necessarily associated with compositions, temperature and pressure, etc). The 2D and 3D
different depositional and diagenetic events. They could be modeling results were consistent with regional and reservoir
associated with the same diagenetic events and in fact, they trends in dolomite/anhydrite distribution published in the
could well be interconnected. The distribution of primary literature and provided a better understanding of key
anhydrite precipitation (bedded anhydrite) and in particular the diagenesis processes controlling reservoir quality in
anhydrite cement can be complex. For example, the anhydrite dolomitized carbonate reservoirs. The modeling results,
cement can be located either at the dolomite front or at close together with well-based observations, generated predictive
approximity to the brine source and may subject to diagenetic rules that can be used to guide porosity and
remobilization by subsequent reflux events. The predicted permeability distribution in 3D reservoir connectivity models.
high-K dolomites were likely to be distributed at the dolomite-
limestone transition zone in a hypersalinity reflux The key conclusions and new insights from the reactive
environment. They tend to be away from the source of the transport model include:
hypersalinity brine so little or no anhydrite occurred. They are
represented by a complex 3D interconnected underground 1. Reactive Transport Models have the capability to simulate
network that could significantly affect reservoir fluid flow. the process of brine reflux dolomitization and the resultant
distribution of dolomite, anhydrite and petrophysical
ASSUMPTIONS AND LIMITATIONS properties in 2 and 3D.

The validity of a reactive transport model depends on how the 2. Sensitivity analyses revealed the hierarchical controls of
(geologically-based) conceptual model is numerically natural variables (salinity, initial lithology, porosity,
represented, how the results are affected by the assumptions permeability and heterogeneity, temperature, and residence
and limitations [11-16, 30-35]. time) on the distribution and evolution of the reflux system.

Our 2D and 3D reflux models only predict reservoir quality 3. Simulations show the propagation of dolomite fronts by
after early reflux diagenesis. The present day reservoir has fingering generating patterns of dolomite consistent with those
gone through additional alteration including compaction and observed in the limited number of outcrop studies.
possible late burial diagenesis such as hydrothermal
dolomitization. Burial alteration likely affects an early 4. Dolomite “fingers” tend to connect to the “main body” in a
dolomite generation, dolomite textures are not incompletely single reflux episode and are rarely isolated “pods”.
altered, and grainstones with pre-existing
porosity/permeability are most likely affected [2, 36-37]. It is 5. The salinity of the reflux brine controls anhydrite primary
likely that “high permeability dolograinstones” are ultimately precipitation and cementation, leading to the formation of
the product of burial alteration (e.g., recrystallization) important diagenetic features such as bedded anhydrite and
superimposed on early dolomites. anhydrite cement, with the latter subject to remobilization.

Uncertainty in the quantitative description of some key 6. Dolomite “sweet spots” are located close to the dolomite
parameters, particularly initial porosity and permeability, front, have a variable thickness and are generally well
limits the precision of reservoir quality predictions using connected in 3D.
reactive transport models [11-16]. Incorporating different
scales of permeability heterogeneity can result in predictions The predicted distribution of dolomite (main body and fingers)
of diagenetic geobodies with a greater resemblance to those anhydrite (bedded anhydrite and anhydrite cement), and high-
observed in nature [11-12]. Another key uncertainty is the K dolomite provided important guidance to develop 3D
specified porosity to permeability feedback relationships for reservoir connectivity models with different scenarios for
different mineral reactions in the variety of carbonate sediment reservoir simulation studies. Future case studies will utilize
types modeled [11]. Scaling up permeability to “effective” well and field data to test the reactive transport models.
values for the dimensions of simulation grids is a further
challenge, particularly when the rocks are fractured [38-39]. In ACKNOWLEDGEMENTS
addition to rock properties, pore fluid composition is also a
significant control on the pattern and magnitude of We thank contributions from ExxonMobil colleagues Sean
groundwater flow and associated mineral reactions [11-12]. Guidry, Jeremy Jameson, Jon Kauffman, Bob Alway, Scott
Parker, Mike Weaver, and Jim Anderson. We also thank Drs.
CONCLUSIONS Tianfu Xu and Karsten Pruess from Lawrence Berkeley
National Laboratory for their generous support in using
We utilized reactive transport models (RTM) to investigate ToughReact program. ExxonMobil’s permission to publish
key controls on reflux dolomitization and reservoir quality. this paper is greatly appreciated.
IPTC 11199 9

APPENDICES: SOFTWARE DESCRIPTION REFERENCES

X2T (University of Illinois) 1. Power, R.W. 1962. Arabian Upper Jurassic carbonate
reservoir rocks. In HAM, W. E. (ed.). Classification of
The X2t program models couples transport and reaction in carbonate rocks. Memoir of the American Association of
geochemical systems open to groundwater flow in one or two Petroleum Geologists, 1, 279 pp.
dimensions. The package includes the full scope of 2. Cantrell, D. Swart, P. and Hagerty, R. 2004. Genesis and
geochemical modeling features of The Geochemist's characterization of dolomite Arab-D reservoir, Ghawar Field,
Workbench¨ (GWB) as well as complete capabilities for Saudi Arabia: Geo-Arabia, v. 9, p. 11–36.
modeling mass transport and heat transfer in one radial or one 3. Insalaco, E,, Virgone, A., Courme, B., Gaillot, J., Kamali,
or two linear dimensions. X2t program can be obtained from M., Moallemi, A., Lotfpour, M. and Monibi, S. 2006. Title:
RockWare Inc. (www.rockware.com) Upper Dalan Member and Kangan Formation between the
Zagros Mountains and offshore fars, Iran: depositional
Tough2/ToughReact (Lawrence Berkeley National system, biostratigraphy and stratigraphic architecture.
Laboratory) Geoarabia, 11 (2): p. 75-176.
4. Wilson, E. N., Hardie, L.A. and Phillips, O.M. 1990.
Tough2 (transport of unsaturated groundwater and heat) is a Dolomitization front geometry, fluid flow patterns, and the
numerical simulator for nonisothermal flows of origin of dolomite: the Triassic Latemar Buildup, northern
multicomponent, multiphase fluids in one, two, and three- Italy: American Journal of Science, v. 290, p. 741-796.
dimensional porous and fractured media. Tough2 uses an 5. Posamentier, H.W., Allen, G.P., James, D.P. and Tesson, M.
integral finite difference (IFD) method for space 1992. Forced regressions in a sequence stratigraphic
discretization, and first-order fully implicit time differencing. framework: Concepts, examples, and exploration
A choice of a sparse direct solver or various preconditioned significance: American Association of Petroleum Geologists
conjugate gradient algorithms is available for linear equation Bulletin, v. 76, p. 1687-1709.
solution. The program provides options for specifying 6. Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M.,
injection or withdrawal of heat and fluids. Double-porosity, Vail, P.R., Sarg, J.F., Loutit, T.S. and Hardenbol, J. 1988. An
dual-permeability, and multiple interacting continua (MINC) overview of the fundamentals of sequence stratigraphy and
methods are available for modeling flow in fractured porous key definitions. In C.K. Wilgus, B.S. Hastings, C.G.St.C.
media. Kendall, H.W. Posamentier, C.A. Ross, J.C. Van Wagoner,
eds., Sea-level changes: an integrated approach. Society of
ToughReact is a non-isothermal reactive geochemical Economic Paleontologists and Mineralogists Special
transport model by introducing reactive geochemistry into the Publication No. 42, p. 39-45.
framework of the existing Tough2 program. ToughReact uses 7. Yao, Q. and Demicco, R.V. 1997. Dolomitization of the
a sequential iteration approach, which solves the transport and Cambrian-Ordovician carbonate platform, southern Canadian
the reaction equations separately. An implicit time-weighting Rocky Mountains: dolomite front geometry, fluid inclusion
scheme is used for individual component of flow, transport, geochemistry, isotopic signature, and hydrogeological
and geochemical reaction. The chemical transport equations modelling studies: American Journal of Science, v. 297, p.
are solved independently for each component, whereas the 892-938.
reaction equations are solved on a grid block basis using 8. Lucia, F.J. and Major, R.P. 1994. Porosity evolution through
Newton-Raphson iteration. An improved equilibrium-kinetics hypersaline reflux dolomitization, in B. Purser, M. Tucker
speciation model for simulating water-rock-gas interaction is and D. Zenger, eds., Dolomites: International Association of
employed. Quasi-stationary approximation and an automatic Sedimentologists Special Publication 21, p. 345-360.
time stepping scheme are implemented in ToughReact. 9. Tillotson, B. 2003. Bed and facies scale selectivity during
late-stage dolomitization: Lower Ordovician El Paso Group,
The methods are applicable to one-, two-, or three-dimensional Franklin Mountains, West Texas. AAPG Annual
geologic domains with physical and chemical heterogeneity. Convention, Salt Lake City, Utah.
Transport of aqueous and gaseous species by advection and 10. Wilson, A. M., Sanford, W.E., Whitaker, F.F. and Smart,
molecular diffusion is considered in both liquid and gas P.L. 2001. Spatial patterns of diagenesis during geothermal
phases. Any number of chemical species in liquid, gas and circulation in carbonate platforms: American Journal of
solid phases can be accommodated. Aqueous complexation, Science, v. 301, p. 727-752.
acid-base, redox, gas dissolution/exsolution, and cation 11. Jones, G.D. and Xiao, Y. 2005. Dolomitization, anhydrite
exchange, are considered under the local equilibrium cementation, and porosity evolution in a reflux system:
assumption. Mineral dissolution and precipitation can proceed Insights from reactive transport models. AAPG Bulletin;
either subject to local equilibrium or kinetic conditions. May 2005; v. 89; no. 5; p. 577-601.
12. Xiao, Y. and Jones, D.J. 2006. Reactive transport modeling
Tough2 and ToughReact programs can be obtained by of carbonate and siliciclastic diagenesis and reservoir quality
contacting Drs. Tianfu Xu and Karsten Pruess from Lawrence prediction. SPE paper #101669, p. 1-10.
Berkeley National Laboratory (www.lbl.gov).
13. Lichtner, P. C. 1996. Continuum formulation of
multicomponent-multiphase reactive transport, in Lichtner,
10 !PTC 11199

P. C., Steefel, C. I. and Oelkers, E. H. (eds.), Reactive 29. Logan, B.W. 1987. The MacLeod evaporite basin, Western
transport in porous media, Reviews in Mineralogy, Mineral Australia: AAPG Memoirs, v. 44, 140 p.
Society of America, v. 34, p. 1-79. 30. Lee. A. 2003. 3-D numerical modeling of freshwater lens on
14. Steefel, C.I, DePaolo, D. and Lichtner, P.C. 2005. Reactive atoll islands. Proceedings, Tough Symposium 2003.
transport modeling: An essential tool and a new research Lawrence Berkeley National Laboratory, Berkeley,
approach for the Earth sciences, Earth and Planetary Science California, 1-7.
Letters 240: p. 539-558. 31. Whitaker, F. F. and Smart, P.L. 1997. Geochemistry of
15. Jones, G.D. and Xiao, Y. 2006. Geothermal convection in the meteoric waters and porosity generation in carbonate islands
Tengiz carbonate platform, Kazakhstan: Reactive transport of the Bahamas. Geofluids II, 2nd International Conference
models of diagenesis and reservoir quality. AAPG Bulletin; on Fluid Evolution, Migration and Interaction in Sedimentary
August 2006; v. 90; no. 8; p. 1251-1272. Basins and Orogenic Belts, University of Belfast, 415-418.
16. Xiao, Y. and Jones, D.J. 2006. Testing carbonate diagenetic 32. Whitaker, F.F., Hague, Y., Smart, P.L, Waltham, D.A. and
paradigms using reactive transport models. AAPG Annual Bosence, D. 1999. Coupled modeling of carbonate diagenesis
Convention, Houston, Programs 11C. and sedimentology: structure and function of a coupled 2-
17. Arvidson, R. S. and Mackenzie, F.T. 1999. The dolomite dimensional diagenetic and sedimentological model of
problem: control of precipitation kinetics by temperature and carbonate platform evolution, Special Publication of the
saturation state: American Journal of Science, v. 299, p. 257- Society of Economic Palaeontologists and Mineralogists 62,
288. p. 69-84.
18. Bethke, C.M. 1996. Geochemical Reaction Modeling, 33. Sanford, W.E., Whitaker, F.F., Smart, P.L. and Jones, G.D.
Concepts and Applications. Oxford University Press, New 1998. Numerical Analysis of seawater circulation in
York, 397 pp. carbonate platforms: I geothermal circulation: American
Journal of Science, v. 298, p. 801-828.
19. Xu, T., Sonnenthal, E.L., Spycher, N. and Pruess, K. 2006.
TOURGHREACT: A simulation program for non-isothermal 34. Kaufman, J.K. 1994. Numerical models of fluid flow in
multiphase reactive geochemical transport in variably carbonate platforms: implications for dolomitization: Journal
saturated geologic media, Computer & Geoscience, of Sedimentary Research, v. A64, p. 128-139.
Lawrence Berkeley National Laboratory. 35. Kohout, F. A., Henry, H.R. and Banks, J.E. 1977. Hydrology
20. Adams, J. E. and Rhodes, M.L. 1960. Dolomitization by related to geothermal conditions of the Floridan Plateau, in
seepage reflux: AAPG Bulletin, v. 44, p. 1912-1920. K. L. Smith and G. M. Griffin (eds) The geothermal nature
of the Floridan Plateau: Florida Department of Natural
21. Sun, S.Q. 1995. Dolomite reservoirs: porosity evolution and
Resources Bureau Geology Special Publication, v. 21, p. 1-
reservoir characteristics: AAPG Bulletin, v. 79, p. 249-257
34.
22. Zenger, D.H., Dunham, J.B. and Ethington, R.L. (eds) 1980.
36. Corbella, M., Ayora, C., Cardellach, E., and Soler, A. 2006.
Concepts and models of dolomitization: SEPM special
Reactive transport modeling and hydrothermal karst genesis:
publication, 28, 320 p.
The example of the Rocabruna barite deposit (Eastern
23. Moore, C.H. 2001. Carbonate reservoirs: porosity evolution Pyrenees). Chemical Geology, Volume 233, Issues 1-2, 30,
and diagenesis in a sequence stratigraphic framework: p. 113-125.
Developments in Sedimentology 55, Elsevier, Netherlands,
37. Garven G., Appold, M.S., Toptygina, V.I. and Hazlett, T.J.
444 p.
1999. Hydrogeologic modelling of the genesis of carbonate-
24. Whitaker, F.F., Smart, P.L. and Jones, G.D. 2004. hosted lead-zinc ores: Hydrogeology Journal v. 7, p. 108-
Dolomitization: from conceptual to numerical models, in C. 126.
Braithwaite, G. Rizzi and G. Darke, eds., The Geometry and
38. Thawer, R., Alhendi, A., Al Mazroui, Y., Boyd, D.,
Petrogenesis of dolomite hydrocarbon reservoirs: Special
Masuzawa, T., Sugawara, Y., Hollis, C. and Lowden, B.
Publication of the Geological Society. Vol. 235, p. 99-139.
2000. Controls on vertical and horizontal flow in a carbonate
25. Jones, G. D. and Rostron, B.J. 2000. Analysis of fluid flow reservoir that impact gasflooding and waterflooding: Society
constraints in regional-scale reflux dolomitization: constant of Petroleum Engineers, SPE87237, 10 p.
versus variable-flux hydrogeological models: Bulletin of
39. Ewing, R.E. 1997. Aspects of upscaling in simulation of flow
Canadian Petroleum Geology, v. 48, p. 230-245.
in porous media: Advances in Water Resources, v. 20, p.
26. Jones, G. D., Smart, P.L., Whitaker, F.F., Rostron, B.J. and 349-358.
Machel, H.G. 2003. Numerical Modeling of reflux
dolomitzation in the Grosmont Platform complex (Upper
Devonian), Western Canada Sedimentary Basin. AAPG
Bulletin, v. 87, p. 1273-1298.
27. Enos, P. and Sawatsky, L.H. 1981. Pore networks in
Holocene carbonate sediments: Journal of Sedimentary
Petrology, v. 51, p. 961-985.
28. Melim, L.A., Westphal, H., Swart, P.K., Eberli, G.P. and
Munnecke, A. 2002. Questioning carbonate diagenetic
paradigms: evidence from the Neogene of the Bahamas.
Marine Geology 185, p. 27 - 53.

You might also like