Sie sind auf Seite 1von 124

Chapter Twelve: Effects of Reservoir Depletion

Topics

• Stress Changes in Depleting Reservoirs


- Stress Path
- Production-Induced Faulting
- Production-Induced Stress Rotations
• Hydraulic Fracturing, Water Flooding and Re-Frac’ing
• Deformation and Stress Changes Inside Depleting
Reservoirs –
−Deformation Analysis in Reservoir Space (DARS)
−Compaction Drive
• Deformation and Stress Changes Outside Depleting
Reservoirs
• Depletion, Compaction and Subsidence
Production-Induced Stress Changes and Faulting

Normal Normal
Faulting Reverse Faulting Faulting

Figure 12.1 – pg. 380 (After Segall, 1989)


Using
Usinginstantaneous
instantaneousapplication
applicationofofforce
forceand
andpressure
pressurewith
withno
nolateral
lateralstrain:
strain:
⎛ ν ⎞ ⎛ ν ⎞
S Hor = ⎜ ⎟ v
( S ) + αP ⎜ 1 − ⎟
⎝ 1 −ν ⎠ ⎝ 1 − ν ⎠
Takethe
Take thederivative
derivativeof
ofboth
bothsides
sidesand
andsimplify
simplify

ΔS Hor = α
(1 − 2ν ) ΔP α = 1−
Kb
(1 − ν ) p Kg
2
if ν = 0.25, α = 1 ΔS Hor = ΔPp
3
ΔS Hor
Stress Path is defined as: A=
ΔPP

Equations 12.1-12.3 – pg. 381


Poro-Elastic Effect of a Sudden Change of Pp (ΔPp)

Stress change immediately above reservoir (Segall and Fitzgerald, 1996):

ΔS h ⎛ 1 − 2ν ⎞⎛ π ⎞⎛ h ⎞
=α⎜ ⎟⎜ ⎟⎜ ⎟
ΔPp ⎝ 1 − ν ⎠⎝ 4 ⎠⎝ L ⎠

Poroelastic stress change within reservoir:

ΔS h ⎛ 1 − 2ν ⎞
= α⎜ ⎟
ΔPp ⎝ 1 −ν ⎠
Stress Path Vs. Biot Coefficient & Poisson’s Ratio

Elastic Solution:
A=α(1-2ν)/(1-ν)
Figure 12.2 – pg. 382
Gulf of Mexico Field X

Figure 12.3 a – pg. 383


Pp (ps i)

0
10
20
30
40
50
60
70
80
90
Feb-82

Nov-84

A ug-87

May-90
Pp
Pp

Jan-93
Field X

Figure 12.3 b – pg. 383


Oct-95

Jul-98
S3

A pr-01
S3

Jan-04
Production Induced Faulting in Normal Faulting Areas

[ SV − ( Pp − ΔPp )]
f (μ ) =
[( Sh min − ΔSh min ) − ( Pp − ΔPp )]

where f ( μ ) = ( μ 2 + 1 + μ ) 2

SV − Pp ⎡ ΔSh min − ΔPp ⎤ ΔPp


= ⎢1 − ⎥ f (μ) −
Sh min − Pp ⎢⎣ Sh min − Pp ⎥⎦ Sh min − Pp

Equations 12.4 & 12.5 – pg. 385


Production Induced Faulting in Normal Faulting Areas

When normal faults are in frictional equilibrium:

⎡ ΔSh min − ΔPp ⎤ ΔPp


f (μ ) = f (μ ) − ⎢ ⎥ f (μ) −
⎢⎣ Sh min − Pp ⎥⎦ Sh min − Pp

⎡ ΔSh min − ΔPp ⎤ ΔPp ⎡ ΔSh min − ΔPp ⎤ 1


⎢ ⎥ f (μ) = − ⎢ ⎥=−
⎢⎣ Sh min − Pp ⎥⎦ Sh min − Pp ⎢⎣ ΔPp ⎥⎦ f (μ )

ΔSh min 1
if A = A* =
ΔPp ( μ 2 + 1 + μ )2
Equation 12.6 – pg. 385
Figure 12.4 a,b – pg. 383
Production-Induced Normal Faulting in the
Valhall Chalk Reservoir, North Sea
NNW-SSE Regional Seismic Section
Valhall Stress Evolution with Production

55
1998

50 SV
1996

45
Depletion 1994

40
1992
Stress [MPa]

35 Normal
1990
Faulting
30
1988
25 μ = 0.9
μ = 0.6 CREST 1986
20
FLANK
1984
15

1982
10
10 15 20 25 30 35 40 45 50 55
Pore Pressure [MPa]
(after Zoback & Zinke, 2001)
Normal Faulting in Crest of the Structure (Natural State)
Stress Path Induces Normal Faulting on the Flanks
(Natural State is a Near-Isotropic Stress State)

Depletion causes a change from a stable stress


regime to a normal faulting regime on the flanks
Valhall Seismology

(after Zoback & Zinke, 2001)


Ekofisk Field
Stress Rotation with Depletion?
Arcabuz Field, Northeast Mexico

Tight gas reservoir


high porosity 10-15%
low perm 0.01 to 0.1 mD

Wilcox Sands
Eocene-Paleocene, shallow marine sands
interbedded with shale

Requires hydrofracing to stimulate


production

Vázquez et al.,
TLE 1997 Fault bounded and
compartamentalized by N-S trending
Setting of the Arcabuz-Culebra
Gas Field Syn-depositional growth faults
region undergoing ESE extension
Arcabuz Field, Northeast Mexico

Arcabuz Field after depletion:


Initial stress and Pp conditions:
aziSHmax ranges from 15° to 148°

aziSH = 21° from failure and


hydrofracs
Sv = 1.07 psi/ft
SHmax slightly > Sv
Shmin slightly under 1 psi/ft at a
maximum from hydrofracs
SHmax – Shmin ~0.2 psi/ft
Pp highly variable, but overall
overpressure with 0.9 psi/ft max

Pp reduction estimated to be anywhere


from 0.09 to 0.4 psi/ft
Stress Rotation with Depletion?

Figure 12.6 c,d – pg. 391


Stress Rotation with Depletion

Figure 12.6 a,b – pg. 391


Scott Field North Sea Quad 15
(Yale et al. 1994)
Depletion in a Laterally Extensive Reservoir Will
Not Cause Stress Orientations to Rotate

⎛ ν ⎞ ⎛ ν ⎞
Using instantaneous application of force and pressure with no lateral strain:
S Hor =⎜ ⎟( Sv ) + αP ⎜1 − ⎟
⎝ 1 −ν ⎠ ⎝ 1 −ν ⎠

ΔS Hor = α
(1 − 2ν )
Take the derivative of both sides and simplify

ΔPp α = 1−
Kb
(1 − ν ) Kg
2
if ν = 0.25, α = 1 ΔS Hor = ΔPp
3
ΔS Hor
Stress Path: A=
ΔPP
Geometry
After depletion

PORE PRESSURE
CHANGES BY ΔPp

Figure 12.7 – pg. 392


Perturbed total stresses

And the new total stresses at the fault in the ORIGINAL


coordinate system are therefore:

AΔPp
S x = S Hmax + (1 − cos 2θ )
2

AΔPp
S y = S hmin + (1 + cos 2θ ) shear means
2
these aren’t
AΔPp principal
τ xy = sin 2θ stresses
2
anymore!
SHmax rotation

We need to rotate the coordinate system (and therefore SHmax!)


through an angle, γ, in order to find new principal stress directions:

1 ⎡ 2τ xy ⎤ 1 −1 ⎡ AΔPp sin 2θ ⎤
γ = − tan ⎢
−1
⎥ = − tan ⎢ ⎥
2 ⎣⎢ x
S − S ⎥
y ⎦ 2 (
⎣ Hmax
S − S hmin ) − A Δ Pp cos 2θ ⎦

− ΔPp
q=
Defining q
(S Hmax − S hmin )

1 ⎡ − Aq sin 2θ ⎤
Yields γ = − tan −1 ⎢ ⎥
2 ⎣1 + Aq cos 2θ ⎦

Equations 12.7-12.12 – pg. 393


Arcabuz Field, Northeast Mexico

Arcabuz Field after depletion:


Initial stress and Pp conditions:
aziSHmax ranges from 15° to 148°

aziSH = 21° from failure and


hydrofracs
Sv = 1.07 psi/ft
SHmax slightly > Sv
Shmin slightly under 1 psi/ft at a
maximum from hydrofracs range of q = 0.45 to 2
SHmax – Shmin ~0.2 psi/ft
Pp highly variable, but overall
overpressure with 0.9 psi/ft max

Pp reduction estimated to be anywhere


from 0.09 to 0.4 psi/ft
Generalized results

A=0.67

q=0

q=1.5

q=10

ΔPp
as increases, so does stress rotation
(S Hmax − S hmin )
Stress Reorientation by Differential Completion

Figure 12.8 a,b – pg. 394


Arcabuz Model Results

− ΔPp
90

q=
(S Hmax − S hmin )
A= 0.66667

60
2

observed γ
Rotation γ (deg) of SHmax

30

from -75° to 85°


0
0.45

-30

-60

-90
-90 -60 -30 0 30 60 90
Angle θ (deg) from original SH to fault
Arcabuz results

90
84
A= 0.66667

60

40
Rotation γ (deg) of SHmax

30

21

0
-7

-30

-54
-60

-75
-90
-90 -60 -30 0 30 60 90
Angle θ (deg) from original SH to fault
Scott Field Results

A = 2/3
q=2
good fit, but sometimes
overpredicts θ
compartments might
require different q’s in
different parts of the
field?
Water Floods and Hydraulic Fracturing

• Many water floods result in pervasive hydraulic fracturing


of the reservoir. If this is not taken into proper account, the
water flood can be ineffective.

• Limiting the vertical growth of hydrofracs requires


knowledge of the least principal stress. This can be
measured, or estimated (like SHmax) from knowledge of
wellbore failure in inclined holes.

• Under some circumstances, production-induced stress


changes can change stress directions. This can make
refracturing of wells possible.

• Hydraulic fracturing (net) pressures in deviated wells can


be minimized by drilling in the appropriate direction.
Step Rate
Extended Leak Off Test
(or Mini-Frac)
Poor Sweep Efficiency
Excellent Sweep Efficiency
Stacked and Compartmentalized Reservoirs
Mountaineer Site Geomechanical Analysis: Results

Regional Direction of SHmax

N World Stress Map, 2004

Lithology

Strike Slip
Injection
Zones
Stress Regime
Application of Results from Geomechanical Analysis
Lower Shmin magnitude in the Rose Run injection zone could
be beneficial to it’s sequestration potential

Motivations: Can I quantify the benefit of hydraulically fracturing the


injection zone? Will this be a worthwhile technique in future projects?
Stress Evolution with Production:
Data from Northeastern Mexico
Pore Pressure Change Affects Fracture Containment

Original closure stress profile; Altered closure stress profile;


original fracture dimensions refrac fracture dimensions

Shale

Sand Depleted

Shale
Re-Fracture Treatment
Results in Longer
Fracture

Original Fracture Grows


Upward and is Short
Do Stress Rotations Caused by Depletion in Faulted Reservoirs
Enhance the Effectiveness of Repeated Hydraulic Fracturing?

Refracturing

Siebrits et al., Refracture reorientation enhances gas


production in Barnett Shale tight gas wells, SPE 2000

US Patent Application
Predicting Changes in Hydrofrac Orientation in Depleting Oil and Gas Reservoirs
by Mark D. Zoback, Amy D.F. Day-Lewis, and Sangmin Kim, Stanford University
Do Stress Rotations Caused by Depletion in Faulted Reservoirs
Enhance the Effectiveness of Repeated Hydraulic Fracturing?

Refracturing

Dozier et al., Refracturing works, Oilfield Review 2003

Siebrits et al., Refracture reorientation enhances gas


production in Barnett Shale tight gas wells, SPE 2000
Multi-Stage Hydraulic Fracturing

(from Schlumberger)
Do Stress Rotations Caused by Depletion in Faulted Reservoirs
Enhance the Effectiveness of Repeated Hydraulic Fracturing?

Refracturing

Dozier et al., Refracturing works, Oilfield Review 2003

Siebrits et al., Refracture reorientation enhances gas


production in Barnett Shale tight gas wells, SPE 2000
Does the rotation in stress direction allow undepleted
sections of the reservoir to be accessed?

Tight gas reservoir


high porosity 10-15%
low perm 0.01 to 0.1 mD

Figure 12.9 – pg. 395


Minimizing Hydraulic Fracturing “Net” Pressures

Problem
Deviated and horizontal wells can be difficult to
hydraulic fracture because small, inclined
fractures form at the wellbore wall

Solution
In open holes, use directional drilling to
minimize fracture initiation and link-up
pressures

In cased holes, use oriented perfs so that the


fractures propagate easily in the direction normal
to the least principal stress
Inclined Tensile Cracks – Geothermal Well, Japan
Occurrence of Drilling-Induced Tensile Fractures

Normal Faulting

Hypothetical Stress
States, Hydrostatic
Pore Pressure

Normal/Strike-
Slip Faulting
Modeling Inclined Tensile Fractures
Link-Up of Inclined Fractures to
Form Larger Fractures

90

80 The criterion for link-up is


70 expressed as:
ω f ≤ ωcrit
60
ω f , deg

50 No Link-up
40
where,
ω f = ω crit
30
⎛ ⎛ ΔS ⎞ −0.72 ⎞

= sin ⎜⎜ 0.57⎜ ⎟ ⎟⎟
−1
20
ω crit
10 Link-up ⎝ ⎝ ΔP ⎠ ⎠
0
0 1 2 3 4 5 and
Δ S/ Δ P

Higher Pw ΔS = Spara – Snorm


ΔP = Pw – Snorm
Critical fracture angle controls the link-up
of inclined fractures near the wellbore to
form larger fractures that can propagate
away from the wellbore.
Modeling Hydrofractures

Hydrofrac Initiation Hydrofrac Link-Up


Deformation Analysis in Reservoir Space
(DARS)

To understand the deformation mechanisms of a


producing reservoir utilizing relatively simple
laboratory tests and in situ measurements

DARS is a formalism for estimating the evolution of


porosity, permeability and the potential for induced
normal faulting in a producing reservoir
GOM Turbidites Porosity Reduction
GOM Turbidites Permeability Reduction

Permeability Change κi/κ

0.1
0.85 0.9 0.95 1
Porosity Change φi/φ
Deformation of Navajo Sand, Moab Fm, Utah
Shear Enhanced Compaction (End Cap)

300 Adamswiller (W97)


Berea (W97)
Boise-2 (W97)
Sv-Sh Darley Dale (W97)
Rothbach-1 (W97)
(MPa) Rothbach-2 (W97)
Kayenta (W97)
Navajo (D73)
200 Kayenta (D73)
15% Cutler (D97)
Adamswiller (W97)
Berea (W97)
Boise (W97)
21% Darley Dale (W97)
Rothbach-2 (W97)
21% Kayenta (W97)
100 Berea (J&T79)
20% Bad Durck (S98)

23% Castlegate (B&J98)


Berea (H63)
Galesville (B81)
Berea (K91)
35%
Vosges (F98)
Red Wildmoor (Pap00)
0
0 100 200 300 400

((Sh+SH+Sv)/3)-Pp (MPa)
Shear Enhanced Compaction (End Cap)
p=
1
3
(σ 1 + σ 2 + σ 3 ) q=
1
2
[
(σ 1 − σ 2 )2 − (σ 1 − σ 3 )2 − (σ 2 − σ 3 )2 ]
Cam-Clay model: M 2 p 2 − M 2 p 0 p + q 2 = 0
DARS

Shmin (MPa)
q (MPa)

Lab Space

Reservoir Space

p (MPa)

Pp (MPa)
DARS

Shmin (MPa)
q (MPa)

Lab Space

Reservoir Space

p (MPa)

Pp (MPa)
Case Study: Gulf of Mexico
Gulf of Mexico Field X
Pp (ps i)

0
10
20
30
40
50
60
70
80
90
Feb-82

Nov-84

A ug-87

May-90
Pp
Pp

Jan-93
Field X

Oct-95

Jul-98
S3

A pr-01
S3

Jan-04
Hydrostatic Compression Tests
Initial porosity
26.5%
Kozeny-Carmen Relationship

∂P k
Darcy’s Law: Q=− A
η ∂x
π 4 ∂P
Laminar flow through a circular pipe: Q=− R
8η ∂x
Effective permeability for the pipe:

⎛ πR 2 ⎞ R 2 Bφ 3 Bφ 3 d 2
k = ⎜⎜ ⎟⎟ = 2 2 =
⎝ A ⎠ 8 τ S τ

Modified Kozeny-Carmen Relationship (Mavko and Nur, 1997):

k=B
(φ − φc )3 d2
k (φ − φc ) (1 + φc − φi )
=
3 2

(1 + φc − φ )2 ki (φi − φc )3 (1 + φc − φ )2
Equations 12.19 & 12.20 – pg. 405-406
Figure 12.13 a – pg. 404
Louisiana
Texas

Field Z

- Pliocene to Miocene Deep water turbidite sand


- initial porosity: ~30 %
- initial permeability: 300 to 2500 md
- initial pressure: ~82 MPa
- 1’ of subsidence at platform
Pressure Data atDataSand
Pressure at Sand U
U ofof
FieldField
Z Z
12000

10000

8000
Pressure (psi)
Pressure (psi)

6000

Well 1

4000
Well 2
Well 3
Well 4
2000 Well 5
Well 6
Well 7
0
5/15/1996

12/1/1996

6/19/1997

1/5/1998

7/24/1998

2/9/1999

8/28/1999

3/15/2000

10/1/2000

4/19/2001
Date

Date
Permeability Reduction
Perm eability Data at Sand U of Field Z

500

450
Well 1
400 Well 2
Well 2
350
Well 3
Well 3
Permeability (mD)

300

250

Well 1
200

150

100

50

Production Tests
Permeability Reduction Vs. Depletion
Perm eability Change Vs Depletion

1.2
Permeability change (k/k0)

Well 1
0.8

Well 2
0.6

0.4
Well 3

0.2

0
0 500 1000 1500 2000 2500 3000 3500

D e p l e t i o n ( p si )
Depletion (psi)
200
In-situ permeabilty Well A
150 measurements Reservoir Quality: Moderate
Permeability

100

50 Predicted permeability from


in-situ stress measurements
0
0 10 20 30 40 50 60 70 80 90 100
Depletion (MPa)
600
Well B
Reservoir Quality: Good
Permeability

400 Lower-bound of permeability loss

200
Upper-bound of permeability loss

0
0 10 20 30 40 50 60 70 80 90 100
Depletion (MPa)
600
Well C
Reservoir Quality: Good
Permeability

400

200

0
0 10 20 30 40 50 60 70 80 90 100
Depletion (MPa)

Figure 12.14 – pg. 408


Summary

DARS is a formalism based on simple laboratory


experiments and in-situ pressure and stress
measurements to study the evolution of a
producing reservoir

DARS does not required the assumption of


poroelasticity
Conceptual Reservoir

•Elliptical reservoir at 16300 ft


depth with single well at centre
•Reservoir dimensions – 6300 x
3150 x 70 ft, grid – 50 x 50 x 1
•Average permeability – 350 md,
φinit – 30%
•Oil flow, little/no water influx, no
injection
•IP – 10 MSTB/d, min. BHP - 1000
psi, Econ. Limit – 100 STB/d
•Ran for maximum time of 8000
days
Simulation Result - Recovery
30
Figure 12.15 – pg. 410
Compaction drive
25
Cum. Oil, MMSTB

20
Compaction drive with
permeability change

15

10

5
Elastic strain only
(Constant compressibility)
0
0 2000 4000 6000 8000 10000
days
•Both conventional cR and no-perm-change curves reached economic limit
•Conventional constant compressibility underestimates the recovery for this case
Field X in GOM

C f = A(σ lab − B )C + D
C f = Δφ Δp
Equation 12.13 – pg. 399

Figure 12.11 – pg. 399


Field X in GOM

Figure 12.16 – pg. 411


Geertsma Model
For a circular reservoir, surface displacements are:

⎧u (r ,0) = −2C (1 −ν )ΔpHR ∞ e − Dα J (αR )J (αr )dα


⎪ z m ∫0 1 0
⎨ ∞
⎪ur (r ,0 ) = 2Cm (1 −ν )ΔpHR ∫ e − Dα J1 (αR )J1 (αr )dα
⎩ 0

Assuming R>>H, total


reduction in reservoir
height:

ΔH = ∫ Cm ( z )Δp( z )dz
H

0
Figure 12.17 a,b – pg. 414
Compaction Model

We only know the approximate geometry of


the reservoir and the amount of depletion.
We need a constitutive law to predict the
amount of compaction (ΔH).

ΔH from creep
compaction law
Subsidence in the Louisiana Coastal Zone due to
Hydrocarbon Production
New Orleans after Hurricane Katrina, August 2005

AP photo
National Geographic, October 2004
Current and Projected Regional Land Loss
Vulnerable ecosystem

One of the largest habitats for


migratory waterfowl in the world

79 plant and animal species in


wetlands listed as threatened or
endangered

Commercial fisheries and aquaculture


are expected to exceed $37B by 2050

Americaswetland.com
Vulnerable energy infrastructure

Over 80% of offshore


oil and gas supply
travels through
Louisiana’s wetlands.

Americaswetland.com
Mechanisms of Coastal Land Loss

Levees and Canals Herbivory


Storms Sea Level Rise
Interior Marsh Losses Subsidence
Edge Erosion

Photo Courtesy of Denise J. Reed


Mechanisms of Subsidence

Modified after Penland et al., 1989


Wetland loss and subsidence due to fluid withdrawal and faulting

USGS Fact Sheet FS-091-01


Oil and gas fields are
pervasive through
the region of high
rates of land loss.

Land Loss 1932-2050


Land Gain 1932-2050
Fluid Production and Wetland Loss: Louisiana Coastal Zone

Modified from Morton et al., 2005


Study Area: LaFourche Parish
Bourg Sand

Approximate surface location of Golden Meadow Fault

Approximate sub-surface location


Approximate surface location of
Golden Meadow Fault

Golden Meadow Fault

Pelican Sand

Bourg Sand
Duval Sand

Exposito Sand

Duval Sand

1993 Relevel line


(Station)
Figure 12.18 a,b – pg. 416
Subsidence Model – All Sands, Elastic-Viscoplastic Rheology
(mm) (mm)

(mm)
Compaction and Subsidence (All Reservoirs)

Approximate location of
Golden Meadow Fault
Vertical Elevation Changes
Approximate Surface Location of
the Golden Meadow Fault

5
Subsidence (cm)

10

15

20

0 6 12
Distance from Station U (km)
Best fit to Geertsma Epoch 1: 1965-1982
(After Morton et al., 2002)
Time-dependent compaction of shales

↓ pp

↓ pp

Modified after Galloway et al., 1999


Also seen in the Central Valley of CA

From USGS Professional Paper 1401-A, "Ground water in the Central Valley, California- A
summary report“ Photo by Dick Ireland, USGS, 1977
Comparison of subsidence rates in 3 modeled epochs

Coast Inland
Regional Implications

A 2006 “Conference of Experts” convened by AGU to integrate science into the


decision-making process for re-building the Gulf Coast following the hurricanes of 2005
recommended:

development of a landscape model which can accurately predict future integrated effects of
subsidence and sediment accretion

improved models of fault movement, petroleum extraction, and water pumping to predict the future
extent of subsidence processes

include patterns and rates of subsidence in planning for the use of water and sediment to nourish
and establish wetlands

http://www.agu.org/report/hurricanes

Das könnte Ihnen auch gefallen