You are on page 1of 16

SPE 143378

Modelling of Depletion-Induced Microseismic Events by Coupled Reservoir


Simulation: Application to Valhall Field

Xing Zhang, SPE, Schlumberger; Nick Koutsabeloulis, SPE, Schlumberger; Tron Kristiansen, BP Norge AS; Kes
Heffer, Reservoir Dynamics Limited; Ian Main and John Greenhough, The University of Edinburgh; Assef
Mohamad Hussein, Schlumberger

C Copyright 2011, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE EUROPEC/EAGE Annual Conference and Exhibition held in Vienna, Austria, 23–26 May 2011.

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 been
reviewed 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, its
officers, 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 to
reproduce 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
Reservoir depletion/injection results in changes in effective stress, which may cause formation failure, propagation of existing
fractures/faults and initialisation of new fractures/faults. A flow-deformation coupled reservoir geomechanical modelling
approach has been applied to compute the change in reservoir effective stress and associated propagation of fractures and
reactivation of faults. The computed inelastic strain change was used to predict the magnitudes of microseismic emissions
associated with fault reactivations. The moment magnitude of microseismic events, consistent with the Richter scale, can be
computed from the scalar value of their moment tensor. In a discontinuous model, the moment tensor can be calculated
directly; in finite element models, the moment tensor is calculated by integrating the change in inelastic shear strain over the
volume of the elements containing the failed fault.
Coupled 3D geomechanical (deformation and fluid flow) simulations for Valhall field were conducted. Well rate and reservoir
pressure histories were used as inputs to simulate reservoir depletion/injection. The 99 faults visible on seismic interpretations
and 22 subregional fracture sets were included in the model. An elasto-inelastic cap model including water-weakening
mechanism was used to model the chalk reservoir; the failure of faults and fractures were simulated by Mohr-Coulomb
criterion. Prior to performing production simulation, preproduction stress modelling was carried out to reach an equilibrium
stress state that was consistent with the in-situ condition, in terms of magnitude and orientation. Then, oil production at 112
wells and water injection at 15 wells for the period from 1982 to 2006 were simulated, during which the simulated
microseismic events were in a good agreement with observations. The computed formation deformation in the reservoir and
overburden was related to the in situ stress and faulting structure and was correlated over long distances. These characteristics
are indicative of a near-critical process (and are also observed in flow rate correlations in the field). The coupled 3D
geomechanical simulation provides a tool for validating modelled reservoir geomechanical effects against specific field data,
and so enhances confidence in the implications predicted for drilling operations and reservoir management.

Introduction
Under most circumstances, an oil or gas field may be described as a system of rock blocks with fluids, partially separated by
faults and fractures. Field evidence has been growing that the interaction between reservoir rock deformation and fluid flow is
an intrinsic component of the system behaviour over the development life of a reservoir. In previous studies (Zhang et al.
2007a, 2007b), three key features—long-range correlation, stress-relationship, and fault-relationship—were observed in
correlations in fluctuations of well flow rates when the stress state in a field was at or close to a critical state. This is because
when field stress state prior to depletion/injection is at or close to a critical state, a relatively small change in effective stress
due to depletion or injection is likely to trigger the sliding of the existing fractures or faults or to initialise new fractures. In this
study, 3D coupled geomechanical deformation and fluid-flow modeling was applied to the Valhall field to understand the
impact of the mechanical-hydraulic interactions on reservoir and overburden deformation, particularly the potential fracture
and fault movements in the overburden induced by reservoir compaction.
Valhall field is located in the southern part of the Norwegian North Sea. The field was discovered in 1975. The water depth in
Valhall field is about 70 m. The reservoir depth varies from 2400 to 2600 m. The primary reservoirs are in the Tor and lower
Hod formations with the Lista formation forming the caprock. Since production started in October 1982, the highly porous,
weak chalk reservoir has undergone significant reservoir compaction, exceeding more than 10 m in some locations in the
reservoir, which has been associated with subsidence of the seafloor exceeding 6 m below the central platform complex
2 SPE 143378

(Kristiansen and Plischke 2010).


In the development of the Valhall 3D geomechanical model, the mechanical properties, in-situ stress condition, and the
information on the existing seismic faults and fracture sets were used, based on which the initial stress state prior to production
was simulated. This model was coupled with flow simulation over 24 years of 112 production wells and 15 water injection
wells, the rates of which were used as inputs, and the resultant changes in reservoir pressure provided perturbations to the
effective stress field.

Development of 3D Geomechanical Model


The 3D geomechanical model was developed using finite element mesh with the implementation of fractures and faults.
Mechanical properties for intact rock, fractures and faults were applied.

Finite Element Mesh.


The finite element mesh was constructed based on the original reservoir flow simulation model (80×44×12) with 8-noded
brick elements. The original reservoir model was refined vertically and was embedded into overburden, sideburden, and
underburden (Fig. 1). The constructed finite element mesh had a total of 256 244 elements (94×58×47).
To reduce the impact of boundary constraints, the overburden was extended up to the seafloor, and the base of the model was
extended down to 64 km. Sixteen layers were added to simulate the overburden formations and nineteen layers were added to
simulate the underburden. In addition, seven rows of elements were added on each side of the original reservoir grids to
construct the sideburden, which was assumed to contain the lateral extension of the reservoir layers.
The top and bottom of the embedded model were assumed to be flat. The shape of the remainder overburden horizons was
based on seismic mapped horizons, whilst the shape of the remainder underburden horizons was graded to match that of the
bottom reservoir horizon shape.

Faults and Fractures.


There were two approximately orthogonal fault sets (Fault pattern 2B) with orientations of about northwest/southeast and
southwest/northeast, as shown in Fig. 2 (left). Faults were modelled by individual elements with suitable material properties.
The faults extended from the reservoir into the overburden and underburden. The methodology adopted within the
geomechanical model is that of an “equivalent” material into which the behaviour of intact rock and discontinuities is lumped.
The equivalent material is thus formulated by ‘smearing out’ the influence of each fault plane throughout the respective
volumes that they occupy.
In this methodology, it is assumed that, within any grid cell that is intersected by at least one fault plane, the rock mass is
composed of the intact rock material and the fault(s). Appropriate constitutive relations (Koutsabeloulis and Zhang 2009) have
been developed for the equivalent material, whose response considers the interaction of all of the constituent components. In
each element intersected by a fault plane, a local system of coordinates (defined by the normal to the fault plane) has been
defined to describe the effective elastic anisotropy of the rock material.
The equivalent material is able to account for both the deformational and failure behaviour in the intact rock and in the faults.
The main difference between these two components is that fault planes permit relative displacements to develop. The
fundamental assumption behind the equivalent-material approach is that these relative movements, when divided by the grid
cell dimensions, can be interpreted as equivalent to continuum strains. The total global strain tensor at each grid cell is
calculated by summing the global strain tensor for the intact rock with those for each of the fault planes. The faults are
simulated with Mohr-Columb failure criterion and the mechanical properties estimated for faults are given in Table 1.

Table 1—Fault mechanical properties


Parameter Value

Normal stiffness (GPa/m) 5


Shear stiffness (GPa/m) 2.5

Friction angle (degrees) 20


Dilation angle (degrees) 10

Cohesion (kPa) 1

Fracture sets were modeled as predefined planes of altered stiffnesses and potential failure with the same equivalent-material
concept. However, they could be taken to represent a range of orientations, sizes, and spacings without specific location in
individual finite elements. The field in this model was divided areally into 22 subregions for the purposes of fracture
characterization, as shown in Fig. 2 (right) which was selected based on a combination of geological and dynamic test data.
SPE 143378 3

Table 2 shows the different sets of fractures and fracture frequency (spacing) interpreted for these subregions. These are
typical average values one finds in the field data (Tjetland et al. 2007). The fracture characteristics were assumed uniform
vertically through the reservoir at each areal location, but did not extend into either overburden or underburden. The
permeability of the fractures varied from 100 to 1000 mD in different fracture sets (Table 3), which is in the order of
magnitude observed in the field. The normal and shear stiffnesses of fractures had the same value and were related to the bulk
and shear modulus of the intact rock, and the fracture spacing also had an effect on the shear and normal stiffness (Table 4).
The friction angle of fracture sets was assumed to be related to the friction angle of the intact rock, which was correlated to the
porosity of the intact rock. The correlation between porosity and friction angle of the intact rock is shown in Fig. 3.

Table 2—Fracture Sets and Fracture Frequency*


Sub- Dip-direction of primary fracture set Dip-direction of secondary Dip-direction of third fracture set
Region fracture set
1 45º L
2 45º L
3 45º L
4 45º L
5 90º L 67.5º L
6 45º M
7 45º L 90º L
8 67.5º M 0º M
9 90º M 0º M
10 67.5º M 135º M
11 90º L 0º L
12 45º H 135º H 0º H
13 45º H 112.5º H 0º H
14 90º M 0º M 135º M
15 45º L
16 135º M 45º M
17 45º M 135º M
18 45º L 0º M
19 135º L
20 45º L
21 45º L
22 45º L
*Vertical fractures were assumed. Where fracture frequency was high (H), 3/m to 30/m; medium (M), 1/m to 3/m; low (L),
<1/m.

Table 3—Permeability of the Fracture Sets


Fracture set Permeability (mD)

Primary Fracture Set 1000


Secondary Fracture Set 300
Third Fracture Set 100

Table 4—Mechanical properties of fractures


Parameter Value
(K, intact rock bulk modulus; G, intact rock shear modulus; s,
fracture spacing)

Normal stiffness (GPa/m) Kn = Ks = (K + 3G/4)/s

Shear stiffness (GPa/m) Kn = Ks = (K + 3G/4)/s

Friction angle (degrees) See Fig. 3


4 SPE 143378

Material Failure Criterion Model and Mechanical Properties of Formations.


Because of the weak nature of chalk with high porosity (0.35 to 0.46), the chalk reservoir was simulated with a chalk model
developed in a joint industry project (Joint Chalk Research) by the Norwegian Geotechnical Institute (NGI). This model
provides different criteria for failure in shear, compression, or tension as determined by the local stress state induced by
depletion and injection. The mechanical properties used are given in Table 5. These properties were estimated as part of this
study and are not necessarily the best parameters for an accurate quantitative prediction, but are accurate enough to capture the
mechanisms investigated in this study. The overburden formations were simulated as elastic material with anisotropic
properties, as shown in Table 6, with embedded faults and fractures.

Table 5—Mechanical Properties for the Chalk Reservoir


Parameter Unit Correlation* Porosity (varies from
0.35 to 0.46)

Young’s modulus MPa 224800 * exp( −11.2φ ) 4460 to 1310

Poisson’s ratio - 0.26 − 0.19φ 0.1935 to 0.1726

Hydrostatic pore strength MPa 500 * exp( −8.5φ ) 25.52 to 10.02

Elliptical cap aspect ratio - φ 0.35 to 0.46

Hardening parameter - (1 + e0 ) =
(1 + e0 ) 3.0 to 4.665

(0.88 − 1.05φ ) (λ − K i )
Transition parameter MPa 5.347 to 0.784
0.1Pc
Cohesion MPa
T * tan (φ ) 0.734 to 0.262

Friction angle Degrees 82.51 − 106.6φ 45.2 to 33.4

Tensile strength MPa


5.0 * exp(− 5.05φ ) 0.729 to 0.398

∗ φ - fractional porosity, Pc – hydrostatic pore collapse strength, T – tensile strength , eo – void


ratio, λ is consolidation index and Ki is inclination of the swelling line

Table 6—Elastic Properties for the Overburden Formations


Parameter Unit Pliocene/ Lower Middle Paleocene
Miocene Miocene Eocene
Young’s Moduli Ex ,Ez
MPa 270 400 650 1500
Shear Modulus Gzx
MPa 108 160 260 577
Young’s Modulus Ey
MPa 165 250 380 1000
Shear Moduli Gxy,Gyz
MPa 45 65 95 350
Poisson’s Ratio
- 0.2 0.2 0.2 0.3

Initial Stress Modelling


In the initial stress modelling, the vertical stress was the overburden weight that was simulated by gravity load. The tectonic
impact was simulated by a specific boundary stress condition. Then, an initial stress modelling was conducted to achieve an
equilibrium state of stress prior to simulating depletion and injection.

Boundary Stress Condition.


Gravity served as the vertical stress and was one of the principal in-situ stresses which was determined by the total overburden
weight. Thus, the maximum and minimum horizontal stresses (SH and Sh) were the other two principal stresses and were
applied on the model boundary with the specific values that were estimated from log data and other measurements. The
vertical total stress (Sv) gradient was 21.5 kPa/m. On the model boundary, principal horizontal stress gradients were applied to
o
simulate the impact of regional tectonic deformation. SH was aligned at N120 E. The total stress Sh/SV ratio was 0.837, and the
total stress SH/SV ratio was 0.93, which produces a total stress Sh/SH ratio of 0.9. The ratios are on the low side of those
SPE 143378 5

reported by Kristiansen (1998) and Kristiansen and Plischke (2010) and thus creates slightly higher shear stresses, but these
are accurate enough to capture the mechanisms investigated in this study.

Simulation of Initial Stresses.


Fig. 4 shows the simulated initial effective maximum horizontal stress, minimum horizontal stress, and vertical stress prior to
the simulation of depletion and injection on a cross-section. Generally, the stresses increased with formation depth because of
gravity loading. The magnitude of the simulated vertical stress was larger than the simulated maximum horizontal stress,
indicating a normal faulting stress regime within the model. The magnitude of the simulated vertical stress was in agreement
with the weight of the overburden, and the magnitudes of the simulated minimum and maximum horizontal stresses were
consistent with the applied boundary stress conditions. The computed inelastic shear strains along faults in the overburden
under the initial stress state are shown in Fig. 5. This indicates that the initial field stress state was at or close to a critical stress
state, particularly along some fault segments. These results indicated that the simulated initial stress state reflected the applied
in-situ conditions.

Coupled Geomechanical Simulation for Depletion and Injection


Based on the simulated initial stress state, the oil production at 112 wells and water injection at 15 wells for the period from
1982 to 2006 were simulated. Geomechanical fluid flow simulations with a consolidation mode were conducted to predict the
changes in stresses during production using coupled geomechanical modeling (Koutsabeloulis and Hope 1998; Koutsabeloulis
and Zhang 2009). During the entire production schedule, the change in pressures, permeability, and effective stresses were
calculated based on the initial pressure, porosity, and permeability. The correlation between pressure and porosity/permeability
can be established based on laboratory rock mechanics tests.
The production/injection operation was simulated on a 1 year by 1 year basis (i.e., the average production rates over a year at
each of the producers and injectors were input). The applied flow rates at the perforations of the wells resulted in a change in
reservoir pressure with time, which induced a change in effective stress within the reservoir. Fig. 6 shows the map view of
pressure change in the top reservoir layer and on a cross section at the end of the simulation period. As a result of pore
pressure change, the effective stresses within the reservoir also varied significantly over time. Fig. 7 shows the effective mean
stress evolution within the upper reservoir through four snapshots during production. The plots show the change of the ratio of
the effective means stresses during production to the effective mean stress prior to production.
As the production activities were dominated by depletion, the reservoir deformation was mainly of compaction. Fig. 8 shows
the induced volumetric strain within the upper reservoir. Significant compressive strains developed, which were strongly
related to the reservoir pressure changes (see Fig. 6). However, the induced strain change was also related to the initial
reservoir rock porosity distribution, mechanical property variation, and the presence of faults.
As the deformation of the chalk reservoir was simulated by a chalk cap model (Koutsabeloulis and Zhang 2009), the reservoir
rock failure, indicated by the development of inelastic strains, could be in shear or compression depending upon the stress state
experienced. Fig. 9 shows the map view of the induced inelastic X-Y shear strains within the upper reservoir with production
time. The inelastic X-Y shear strains indicated the east/west component of reservoir rock shearing in a normal faulting fashion
due to sliding. The induced inelastic shear strains within reservoir increased with production time due to progressive failure of
the reservoir rock. Fig. 10 shows the induced inelastic X-Y shear strains within different reservoir layers at the end of
production. Generally, the induced inelastic shear strains within reservoir decreased with depth. This is because the induced
inelastic shear strains were strongly related to the mechanical property variation and the depletion level. In addition, the
presence of faults had also an influence on the inelastic shear pattern. The inelastic Y-Z shear strains (north/south component)
were significantly smaller than the inelastic X-Y shear strains (east/west component). This is due to an approximately north-
northwest/south-southeast fault strike and a north-northeast orientation of the minimum horizontal stress (Shmin), which
demonstrated that the shear failure of reservoir rock due to depletion was strongly fault and stress related. The strike slip
inelastic component of induced inelastic shear strains (indicated by the Z-X shear strains) was much smaller than the normal
faulting components, particularly the east/west component. This demonstrated again that the reservoir failure due to depletion /
injection activities was strongly related to the imposed in-situ stress regime.
The large degree of reservoir compaction also resulted in significant horizontal movements of the reservoir rock. Fig. 11
shows the computed horizontal north/south and east/west displacements on the reservoir top during production. The east/west
displacements were significantly larger than the north/south displacement due to an approximately north-northwest/south-
southeast fault strike and a north-northeast orientation of the Shmin. This is also consistent with the prediction by Kristiansen
and Plischke (2010), in which the horizontal displacement at seafloor was dominated by the east/west component. The
maximum horizontal displacement at the top or the reservoir was about 2 m, which was also close to the value at seafloor
predicted by Kristiansen and Plischke (2010).
To assess the state of deformation in the overburden associated with fault reactivation, the evolution of the induced inelastic
and total shear strains in the overburden during production is compared, as shown in Fig. 12. It is apparent that the reservoir
compaction resulted in fault reactivation, indicated by the increasing inelastic shear strains along faults with production. The
inelastic shearing along the faults was accompanied by elastic shearing at and around the faults. This can be identified by the
higher values and larger areas of the total shear strains compared to those of the inelastic shear strains. In addition, the
reservoir compaction caused significant deformation in the overburden above the reservoir edges. The deformation was elastic,
6 SPE 143378

but the magnitude was significant. In general, the pattern of fault reactivation in the overburden was strongly related to the
deformation of the underlying reservoir over a long distance.
In this model, the entire overburden was simulated as hydraulically inactive (i.e., the pore pressure within the overburden was
kept constant during the production). However, because of significant deformation caused by the compaction of the reservoir,
the stresses in the overburden were also altered. This change in stresses was due to the mechanical deformation of the
overburden formations, rather than change in pore pressure. In real cases with significant pressure difference between reservoir
and caprock for a long time, the change in pore pressure will also add to the stress change as discussed by Kristiansen and
Plischke (2010). Fig. 13 shows the evolution of the vertical stress and minimum horizontal stress within the overburden, which
is indicated by the change of the ratio of the stress during production to the stress prior to production. Because of the
reactivation of faults that partially extend into the overburden, the stress changes were mainly around the faults. The vertical
stress increased and decreased at different locations because of stress arching. The change in the minimum horizontal stress
was more significant and strongly fault related. The minimum horizontal stress mainly decreased in the central area and
increased above the edge of reservoir basins. The impact of these stress changes on long-term well life and how well design
can be made more efficient in this environment is presented by Kristiansen et al. (2000). The impact on drilling operations and
how computational geomechanics can be used in detailed well planning to reduce drilling risk is presented by Kristiansen and
Flatebø (2009).
The same elasto-inelastic cap model with water-weakening mechanism was also used to model the chalk reservoir. However,
no significant change was computed with and without water-weakening mechanism, in terms of computed reservoir
deformation and deformation-induced fault reactivation. This observation is also in line with the work by Kristiansen and
Plischke (2010), including more recent field data.

Discussion and Conclusion


The moment magnitude of microseismic events, consistent with the Richter scale, can be computed from the scalar value of
their moment tensor through the relationship by Hanks and Kanamori (1979). Moment tensor is the product of the rigidity
modulus, the fault surface area, and the average slip along the fault; the scalar moment is formed through the root-mean-square
of the eigenvectors of the tensor. In a discontinuous model, the moment tensor can be calculated directly (e.g., Hazzard and
Young 2002); in this study (a finite element model), the moment tensor was calculated by integrating the change in inelastic
shear strain over the volume of the elements containing the failed fault.
Microseismic monitoring experiments were performed at both Valhall and Ekofisk (Zoback and Zinke 2002). At Valhall, an
array of seismometers was deployed between June 1 and July 27, 1998 in a vertical section of one of the wells near the crest of
the structure. The seismic array was deployed about 300 m above the reservoir, which recorded the locations of 328 high-
quality microseismic events during the ~7-week monitoring period.
At Valhall, the measured microseismic events in the overburden were recorded within the depth between 2250 and 2400 m
(Kristiansen et al. 2000; Zoback and Zinke 2002). The computed inelastic shear strains in the overburden in this study were
recorded within depths between 2200 and 2400 m. The comparison between the measured microseismic events and the
computed inelastic shear strains is shown in Fig. 14.
The intensity of fault reactivation (indication of microseismic events) varied during the production periods. Relatively strong
fault reactivations were computed during the period between 1997 and 2000, during which microseismic events were recorded
in the overburden. However, this model also computed a stronger period of fault reactivations in the overburden between 2000
and 2003, but a weaker period of fault reactivations in the overburden between 2003 and 2006.
The 3D coupled reservoir simulations have demonstrated that depletion-induced reservoir compaction has a significant impact
on the deformation of the overburden related to stress change, displacements, strains, and microseismic events. This depletion-
related deformation also featured long-range, stress- and fault–related characteristics that are also prevalent in microseismic
data and correlations of flow rate fluctuations (Main et al. 2007). Field test data indicate that passive seismic monitoring has
the potential to provide valuable engineering information for reservoir management, such as well placement planning, well
design, waterflood schedule, hydraulic fracturing monitoring and drill-cuttings re-injection (DCRI), and in more applications
as presented by Kristiansen (2009). The modelling of microseismic events using coupled-reservoir simulation provides a tool
for validating modelled reservoir geomechanical effects against specific field data, and so enhances confidence in the
implications predicted for drilling operations and reservoir management. Even if the modelling presented here did not focus on
the detailed history match to field data at Valhall as presented by Kristiansen and Plischke (2010), the results are accurate
enough to demonstrate the mechanisms that we have investigated and presented in this study.

Acknowledgments
The authors wish to thank BP Norge AS and Hess Norge AS for permission to publish this work. The conclusions and views
presented in this paper are the authors’ and do not necessarily reflect the view of the Valhall license. The authors also would
like to thank our colleagues Bryan Galloway, Jennifer Smith and Suhas Bodwadkar for reviewing the manuscript and
acknowledge Schlumberger for permission to publish this paper.
SPE 143378 7

Nomenclature
Ex = Intact rock Young’s modulus in x-direction, MPa
Ey = Intact rock Young’s modulus in y-direction, MPa
Ez = Intact rock Young’s modulus in z-direction, MPa
G = Intact rock shear modulus, MPa
Gxy = Intact rock shear modulus between x and y directions, MPa
Gyz = Intact rock shear modulus between y and z directions, MPa
Gzx = Intact rock shear modulus between z and x directions, MPa
K = Intact rock bulk modulus, MPa
Ki = Inclination of the swelling line.
Kn = Fracture normal stiffness, GPa/m
Ks = Fracture shear stiffness, GPa/m
Pc = Hydrostatic pore collapse strength, MPa
T = Tensile strength, MPa
eo = Void ratio
s = Fracture spacing, m
φ = Fractional porosity
λ = Consolidation index

References
Hanks, T.H. and Kanamori, H. 1979. A Moment Magnitude Scale, J. Geophys. Res. 84, 2348.
Hazzard, J.F. and Young, R.P. 2002. Moment Tensors and Micromechanical Models, Tectonophys 356, 181–197.
Koutsabeloulis, N.C. and Hope, S.A. 1998. “Coupled” Stress/Fluid/Thermal Multi-Phase Reservoir Simulation Studies Incorporating Rock
Mechanics. Paper SPE/ISRM 47393 presented at the SPE Annual Technical Conference and Exhibition, Trondheim, Norway, 8–10 July.
Koutsabeloulis, N. and Zhang, X. 2009. 3D Reservoir Geomechanical Modeling in Oil/Gas Field Production. Paper SPE 126095 presented at
the SPE Saudi Arabia Section Technical Symposium and Exhibition held in Alkhobar, Saudi Arabia, 9–11 May.
Kristiansen, T.G. 1998. Geomechanical Characterization of the Overburden Above the Compacting Chalk Reservoir at Valhall. Paper
SPE/ISRM 47348 presented at SPE/ISRM Rock Mechanics in Petroleum Engineering, Trondheim, Norway, 8–10 July.
Kristiansen, T.G. 2009 Computational Mechanics Based Workflow and Toolkit to Manage Deformable Reservoir Developments. Paper SPE
124131 presented at the 2009 SPE Offshore Europe Oil & Gas Conference & Exhibition, Aberdeen, UK, 8–11 September 2009.
Kristiansen, T.G., Barkved, O., and Pattillo, P.D. 2000. Use of passive seismic monitoring in well and casing design in the compacting and
subsiding Valhall field, North Sea. Paper SPE 65134 presented at the SPE European Petroleum Conference, Paris, France, 24–25 October.
Kristiansen, T.G. and Flatebø, R.E. 2009. 60 Days Ahead of Schedule—Reducing Drilling Risk at Valhall Using Computational
Geomechanics. Paper SPE/IADC 119509 presented at the SPE/IADC Drilling Conference and Exhibition, Amsterdam, The Netherlands,
17–19 March.
Kristiansen, T.G. and Plischke, B. 2010. History Matched Full Field Geomechanics Model of Valhall Field Including Water Weakening and
Re-Pressurisation. Paper SPE 131505 presented at the SPE EUROPE/EAGE Annual Conference and Exhibition, 14–17 July.
Main, I.G., Li, L., Heffer, K.J., Papasouliotis, O., Leonard, T., Koutsabeloulis, N.C., and Zhang, X. 2007. The Statistical Reservoir Model:
Calibrating faults and fractures, and predicting reservoir response to water flood. In: Structurally Complex Reservoirs, ed. S.J. Jolley, D.
Barr, J.J. Walsh, and R.J. Knipe, Geological Society, London, Special Publications 292, 469–482. doi: 10.1144/SP292.25
Tjetland, G., Kristiansen, T.G., and Buer, K. 2007. Reservoir Management Aspects of Early Waterflood Response after 25 Years of
Depletion in the Valhall Field, Paper IPTC 11276 presented at the International Petroleum Technology Conference, Dubai, U.A.E., 4–6
December 2007.
Zhang, X., Koutsabeloulis, N., and Heffer, K. 2007a. Hydro-mechanical modelling of critically stressed, faulted reservoirs, American
Association of Petroleum Geologists Bulletin 91: 31–50.
Zhang, X., Koutsabeloulis, N., Heffer, K., Main, I. and Li, L. 2007b. Coupled geomechanics-flow modelling at and below a critical stress-
state used to investigate common statistical properties of field production data. In Structurally Complex Reservoirs, ed S. J. Jolley, D. Barr,
J. J. Walsh, and R. J. Knipe, Special Publications, London: Geological Society 292: 453-468.
Zoback, M.D. and Zinke, J.C. 2002. Production-Induced Normal Faulting in the Valhall and Ekofisk Oil Fields, Pure and Applied
Geophysics 159, 403–420.
8 SPE 143378

Fig. 1—Embedded 3D geomechanical model (upper) based on the original reservoir model (lower).

Fig. 2—Implemented 99 seismic faults (left); simulated 22 subregions with different fracture sets (right).
SPE 143378 9

Fig. 3—The friction angle of fracture sets was assumed to be related to the friction angle of the intact rock, which was correlated to
the porosity of the intact rock.

Effective maximum horizontal stress (kPa) Effective minimum horizontal stress (kPa)

Effective vertical stress (kPa)


Fig. 4—Simulated initial effective maximum horizontal, minimum horizontal, and vertical stresses on a cross-section
10 SPE 143378

SH
Sh

X (E)
SH Sh

Z (S)

Fig. 5—Computed inelastic shear strains within an overburden layer at the initial stress conditions that indicated that the faults were
close to or in a critical stress state. SH: Maximum horizontal stress; Sh: Minimum horizontal stress.

Fig. 6—Map view of pressure change (colour scale: 0 to 40 000 kPa) due to depletion and injection within the upper reservoir (upper)
and on a cross-section (lower) at the end of production.
SPE 143378 11

After 6 years (1988) After 12 years (1994)


SH
Sh

SH
Sh

After 18 years (2000) After 24 years (2006)

X (E)

Z (S)

Fig. 7—Effective mean stress evolution within upper reservoir during production. The plots show the change of the ratio of the
effective mean stresses during production to the effective mean stress prior to production. SH: Maximum horizontal stress; Sh:
Minimum horizontal stress.

After 6 years (1988) After 12 years (1994)

After 18 years (2000) After 24 years (2006)

Fig. 8—Volumetric strain development within upper reservoir during production. Compressive volumetric strain increased with
production because of the increase of effective mean stress.
12 SPE 143378

After 6 years (1988) After 12 years (1994)


SH
Sh

SH
Sh

After 18 years (2000) After 24 years (2006)

X (E)

Z (S)

Fig. 9—Induced inelastic X-Y shear strains (colour scale: -0.02 to 0.02) within upper reservoir during production. The inelastic X-Y
shear strains indicated the east/west component of reservoir rock shearing in a normal faulting fashion due to sliding. SH: Maximum
horizontal stress; Sh: Minimum horizontal stress.

Upper Upper middle

Lower middle Lower

Fig. 10—Induced inelastic X-Y shear strains (colour scale: -0.02 to 0.02) within different reservoir layers at the end of production.
SPE 143378 13
North/south displacement East/west displacement

SH

Sh

SH
Sh
After 6 years (1988)

After 12 years (1994)

After 18 years (2000)

X (E)

Z (S) After 24 years (2006)

Fig. 11—Computed horizontal displacements (m) on the reservoir top during production. SH: Maximum horizontal stress; Sh:
Minimum horizontal stress.
14 SPE 143378

Inelastic shear strain Total shear strain

SH

Sh

Sh SH
After 6 years (1988)

After 12 years (1994)

After 18 years (2000)

X (E)

Z (S)
After 24 years (2006)

Fig. 12—Computed inelastic and total shear strains in the overburden during production. SH: Maximum horizontal stress; Sh:
Minimum horizontal stress.
SPE 143378 15

Vertical stress ratio Minimum horizontal stress ratio

SH

Sh

SH
Sh
After 6 years (1988)

After 12 years (1994)

After 18 years (2000)

X (E)

Z (S)
After 24 years (2006)

Fig. 13—Change of the ratio of stress during production to the stress prior to production. SH: Maximum horizontal stress; Sh:
Minimum horizontal stress.
16 SPE 143378

Fig. 14—Measured microseismic events (upper and middle, after Zoback and Zinke 2002; Kristiansen et al. 2000) in the overburden
within the depth between 2250 and 2400 m in 1998 and the computed moment magnitude (lower left) and inelastic shear strains (lower
right) in the overburden within depth between 2200 and 2700 m during the same period, indicating fault-reactivation-induced
microseismic events.

You might also like