You are on page 1of 903

Fracturing Engineering Manual

Preface

The intention of the Fracturing Engineering Manual is to document


Dowell fracturing engineering technology so that it will be uniformly
available to Dowell personnel. The nature and application of the
services, techniques, and associated products used in fracturing
treatment design, execution, and evaluation are included. The manual is
to be instructional in nature. Although the emphasis will be on
application of technology rather than technical details, some technical
details are not always available everywhere and are included.

Every attempt has been made to make the Fracturing Engineering


Manual as up to date as possible, realizing that every manual at the time
of its publication is partially out of date. Each section in the Fracturing
Engineering Manual will be periodically reviewed. Revisions and
additions will be distributed when necessary.

This Preface is hyperlinked to the Master Table of Contents which


hyperlinks to all the sections in the manual. Click on the MASTER
TABLE OF CONTENTS below to jump to that location.

MASTER TABLE OF CONTENTS


FRACTURING ENGINEERING
MANUAL

Schlumberger
Dowell
ITM-1095
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 1 of 33

MASTER TABLE OF CONTENTS


Section 100 Reservoir Evaluation
1 Introductory Summary................................................................................................................ 4
2 Well Performance ...................................................................................................................... 5
2.1 Inflow Performance .............................................................................................................. 6
2.2 Tubing Intake ....................................................................................................................... 8
3 Well Test Interpretation............................................................................................................ 10
3.1 Critical Variables ................................................................................................................ 11
3.2 Flow Regimes .................................................................................................................... 13
3.3 Boundary Effects................................................................................................................ 14
3.4 Diagnostic Plots ................................................................................................................. 15
3.5 Type Curves....................................................................................................................... 18
3.6 Computational System ....................................................................................................... 21
3.7 Steps for Analysis .............................................................................................................. 27
3.8 Example Analysis............................................................................................................... 30
4 Economic Analysis ................................................................................................................... 38
4.1 FracNPV Software ............................................................................................................. 40
5 Application ............................................................................................................................... 48
6 Equation Summary .................................................................................................................. 55
6.1 Oil IPR Equations............................................................................................................... 55
6.1.1 Darcy's Law .............................................................................................................. 55
6.1.2 Vogel Test Data ( Pr ≤ pb ) ........................................................................................ 55
6.1.3 Combination Vogel = Darcy Test Data r ≥ b ...................................................... 55
(P p )
6.1.4 Jones IPR ................................................................................................................. 57
6.2 Gas IPR Equations............................................................................................................. 58
6.2.1 Darcy's Law (Gas) .................................................................................................... 58
6.2.2 Jones' Gas IPR (General Form) ............................................................................... 58
6.3 Backpressure Equation ...................................................................................................... 59
6.4 Transient Period Equations................................................................................................ 60
6.4.1 Time to Pseudosteady State..................................................................................... 60
6.4.2 Oil IPR (Transient) .................................................................................................... 60
6.4.3 Gas IPR (Transient).................................................................................................. 61
6.5 Completion Pressure Drop Equations................................................................................ 61
6.5.1 Gravel-Packed Wells ................................................................................................ 61
6.5.2 Open Perforation Pressure Drop .............................................................................. 63
Section 100 Figures
Fig. 1. Pressure losses in complete systems (after Mach, Proano and Brown)........................... 5
Fig. 2. Location of various nodes (Mach et al., 1981). ................................................................. 6
Fig. 3. Typical IPR curve. ............................................................................................................. 7
Fig. 4. Vogel's composite IPR. ..................................................................................................... 8
Fig. 5. Vertical multiphase flow: How to find the flowing bottomhole pressure. .......................... 9

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 2 of 33

Fig. 6. Tubing intake curve. ..........................................................................................................9


Fig. 7. Deliverability of the producing system. ............................................................................10
Fig. 8. Well log............................................................................................................................12
Fig. 9. Well log............................................................................................................................12
Fig. 10. Radial flow. ....................................................................................................................13
Fig. 11. Linear flow in the formation. ..........................................................................................14
Fig. 12. Bilinear flow. ..................................................................................................................14
Fig. 13. Conditions associated with the boundries. ....................................................................15
Fig. 14. Pressure change and elapsed time to use in a drawdown. ...........................................16
Fig. 15. Log-Log plot. .................................................................................................................17
Fig. 16. Pressure change and elapsed time. ..............................................................................17
Fig. 17. Complete log-log behavior.............................................................................................17
Fig. 18. This is a reproduction from a type curve described in World Oil, (Oct. 1983). ..............19
Fig. 19. Series of pressure, pressure derivative, and specialized plots for common
reservoir features..........................................................................................................20
Fig. 20. Diagnostic log-log plot. ..................................................................................................21
Fig. 21. Two Horner plots. ..........................................................................................................22
Fig. 22. Model-Verified Interpretation. ........................................................................................23
Fig. 23. Conceptual model catalog. ............................................................................................24
Fig. 24. NODAL plot. ..................................................................................................................25
Fig. 25. Sequence simulation. ....................................................................................................25
Fig. 26. Simulated validation. .....................................................................................................26
Fig. 27. PVT plot.........................................................................................................................26
Fig. 28. Matching a diagnostic log-log plot to a type curve.........................................................29
Fig. 29. Matching a diagnostic log-log plot to a type curve.........................................................29
Fig. 30. Log-Log Plot ..................................................................................................................30
Fig. 31. Generated type curve with the log-log diagnostic match...............................................32
Fig. 32. Semilog presentation using a superposition type curve and the data points
from the buildup. ...........................................................................................................33
Fig. 33. Cartesian plot of the simulated pressure and the actual measured data. .....................33
Fig. 34. Decline curve and sensitivity plot. .................................................................................34
Fig. 35. Decline curve and sensitivity plot. .................................................................................34
Fig. 36. Sensitivity plot................................................................................................................35
Fig. 37. Sensitivity plot................................................................................................................35
Fig. 38. Sensitivity plot................................................................................................................36
Fig. 39. Sensitivity plot................................................................................................................36
Fig. 40. Sensitivity plot................................................................................................................37
Fig. 41. Plot of the transient IPR curves with the tubing intake and wellhead
pressure of 875 psi. ......................................................................................................37
Fig. 42. Transient IPR plot for the same tubing, but using different wellhead pressures to
generate the plot shown in Fig. 32................................................................................38
Fig. 43. Conceptual NPV calculation. Case A: revenue is larger than the cost, resulting in a
positive NPV; Case B: revenue is less than the cost, resulting in a negative NPV.......40

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 3 of 33

Fig. 44. Components of the NPV calculation. ............................................................................ 41


Fig. 45. Fracture in a bounded reservoir.................................................................................... 41
Fig. 46. Constant-rate type curve for finite-conductivity fracture  closed square system
(xe/ye = 1). ...................................................................................................................42
Fig. 47. Constant-rate type curve for finite-conductivity fracture  closed rectangular system
(xe/ye = 2). ...................................................................................................................42
Fig. 48. Constant-rate type curve for finite-conductivity fracture  closed rectangular system
(xe/ye = 4). ...................................................................................................................43
Fig. 49. Transient IPR for a fracture well in a closed square reservoir. ..................................... 44
Fig. 50. Tubing intake curve....................................................................................................... 45
Fig. 51. FracNPV analysis for one, two and three years versus fracture length. ....................... 46
Fig. 52. Cumulative production versus fracture length............................................................... 46
Fig. 53. Production decline versus time. .................................................................................... 47
Fig. 54. Cumulative production versus time. .............................................................................. 47
Fig. 55. Well performance tracking form. ................................................................................... 51
Fig. 56. Well performance tracking form. ................................................................................... 53
Section 100 Tables
Table 1. Results of Buildup Test ................................................................................................ 31
Table 2. Maximum Fracture Lengths for Drainage Shapes ....................................................... 43
Section 200 Rock Mechanics
1 Introductory Summary................................................................................................................ 3
2 Lithology..................................................................................................................................... 3
3 Basic Concepts .......................................................................................................................... 9
3.1 Stress Concept .................................................................................................................... 9
3.2 Stress Definition ................................................................................................................. 10
3.3 Strain.................................................................................................................................. 17
3.4 Modulus/Poisson Effect...................................................................................................... 18
3.5 Effective Stress .................................................................................................................. 22
3.6 Failure Criteria ................................................................................................................... 23
4 Measurement of Rock Properties ............................................................................................ 27
4.1 Typical Properties .............................................................................................................. 29
4.2 Toughness ......................................................................................................................... 31
5 Determining in-Situ Stress ....................................................................................................... 35
5.1 Core Tests ......................................................................................................................... 35
5.2 Microfracturing ................................................................................................................... 42
5.3 Pump-In/Flowback Test ..................................................................................................... 44
5.4 Logs ................................................................................................................................... 44
6 Fracture Height (Post).............................................................................................................. 47
6.1 Radioactive Tracer............................................................................................................. 47
6.2 Temperature ...................................................................................................................... 50

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 4 of 33

Section 200 Figures


Fig. 1. Example Litho-Density* log and CNL* Compensated Neutron log. ...................................6
Fig. 2. Example crossplot chart. ...................................................................................................7
Fig. 3. Lithology log. .....................................................................................................................8
Fig. 4. Triaxial stress. .................................................................................................................10
Fig. 5. Randomly oriented plane.................................................................................................11
Fig. 6. Rectangular axis system. ................................................................................................12
Fig. 7. Two-dimensional systems. ..............................................................................................13
Fig. 8. Stress charts....................................................................................................................15
Fig. 9. Strain measurement. .......................................................................................................17
Fig. 10. Typical stress/strain curves. ..........................................................................................18
Fig. 11. Stress strain relationship. ..............................................................................................20
Fig. 12. Compressional force......................................................................................................21
Fig. 13. Principal stresses in a triaxial application. .....................................................................23
Fig. 14. Mohr's circle. .................................................................................................................24
Fig. 15. Mohr failure envelope. ...................................................................................................25
Fig. 16. Typical failure envelope.................................................................................................26
Fig. 17. Dynamic elastic properties. ...........................................................................................27
Fig. 18. Permeability versus confining stress for a fractured rock specimen..............................30
Fig. 19. Stress concentration near the tip of a crack. .................................................................32
Fig. 20. Modified ring. .................................................................................................................33
Fig. 21. Load/displacement curve obtained during a Modified Ring Test...................................33
Fig. 22. Drawing of the LVDT* displacement gauge developed at the Sandia National
Laboratory1....................................................................................................................36
Fig. 23. Typical ASR curve. ........................................................................................................37
Fig. 24. Anelastic strain recovery measurements.......................................................................38
Fig. 25. Gage pattern. ................................................................................................................39
Fig. 26. Typical DSCA plots........................................................................................................40
Fig. 27. Stereoplot. .....................................................................................................................41
Fig. 28. Idealized plot. ................................................................................................................42
Fig. 29. Microfrac test. ................................................................................................................43
Fig. 30. Fracture-height log. .......................................................................................................46
Fig. 31. Two-isotope tracking of a two-stage fracture treatment. ...............................................48
Fig. 32. Two-isotope tracking of a single-stage fracture treatment.............................................49
Fig. 33. Temperature survey with RA tracer...............................................................................51
Section 200 Tables
Table 1. X-Ray Diffraction Analysis ..............................................................................................4
Table 2. Dynamic Versus Static Tests........................................................................................28
Section 300 Fracture Modeling
1 Introductory Summary ................................................................................................................2
2 Concepts ....................................................................................................................................3
2.1 Fundamental Laws ...............................................................................................................4
2.2 Constitutive Laws .................................................................................................................4
2.3 Fracture Propagation............................................................................................................6

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 5 of 33

3 Hydraulic Fracturing Models ...................................................................................................... 9


3.1 Two-Dimensional (2D) ....................................................................................................... 11
3.2 Pseudo Three-Dimensional (P-3D) .................................................................................... 15
3.3 Planar Three-Dimensional (PL-3D).................................................................................... 18
3.4 Fully Three-Dimensional (3D) ............................................................................................ 19
4 Examples ................................................................................................................................. 20
4.1 Case History ...................................................................................................................... 20
4.2 Model Comparisons ........................................................................................................... 28
Section 300 Figures
Fig. 1. Modes of loading............................................................................................................... 7
Fig. 2. Fracture divided into elements.......................................................................................... 9
Fig. 3. Representation of a planar fracture. ............................................................................... 10
Fig. 4. KGD geometry. ............................................................................................................... 11
Fig. 5. PKN geometry................................................................................................................. 12
Fig. 6. 2D and radial Sneddon cracks. ....................................................................................... 13
Fig. 7. Elliptical profile (P-3D)..................................................................................................... 17
Fig. 8. Example grid (PL-3D model)........................................................................................... 18
Fig. 9. Fracture profile (PL-3D model). ...................................................................................... 19
Fig. 10. Permeability, thickness and stress profile. .................................................................... 20
Fig. 11. Computed values for Young's modulus and Poisson's ratio. ........................................ 21
Fig. 12. Profile of bottomhole, casing and tubing pressures. ..................................................... 24
Fig. 13. Pressure match for bottomhole and casing pressure. .................................................. 24
Fig. 14. Fracture profile.............................................................................................................. 25
Fig. 15. Fracture width profile. ................................................................................................... 25
Fig. 16. Match of net pressure for calibration fracture and main fracture. ................................. 26
Fig. 17. Fracture profile.............................................................................................................. 26
Fig. 18. Reservoir model for final history match......................................................................... 28
Fig. 19. TRIFRAC length and width profile................................................................................. 32
Fig. 20. STIMPLAN length and width. ........................................................................................ 32
Fig. 21. FRACPRO length and width profile............................................................................... 33
Fig. 22. GOHFER length and width profile................................................................................. 33
Fig. 23. TERRAFRAC length profile........................................................................................... 34
Fig. 24. STIMPLAN length and width profile. ............................................................................. 34
Fig. 25. MEYER length and width profile. .................................................................................. 35
Fig. 26. Ohio state length profile. ............................................................................................... 35
Section 300 Tables
Table 1. Comparison Of Stress.................................................................................................. 22
Table 2. Permeability and Fluid Loss ......................................................................................... 22
Table 3. Design Information....................................................................................................... 23
Table 4. Fracture Model Comparison Runs ............................................................................... 30
Table 5. Fracture Model Comparison Runs ............................................................................... 31

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 6 of 33

Section 400 Treatment Design


1 Introductory Summary ................................................................................................................3
1.1 References ...........................................................................................................................3
2 Adaptive Design Methodology....................................................................................................6
2.1 Setting Characterization .......................................................................................................7
2.2 Setting Modelization .............................................................................................................8
2.3 Adaptive Treatment Design ................................................................................................10
3 Data Collection .........................................................................................................................10
4 Fracturing Fluid Selection.........................................................................................................12
4.1 Fracturing Fluid Selection Guide ........................................................................................13
4.2 Water Sensitivity of the Reservoir Rock .............................................................................16
4.3 Rheological Properties and Viscosity Requirements ..........................................................17
4.4 Fluid Friction Pressure........................................................................................................18
4.5 Fluid Compatibility with Reservoir Fluid and Rock..............................................................20
4.6 Rheology Selection.............................................................................................................20
4.7 Fracturing Fluid Additive Selection .....................................................................................21
5 Proppant Selection ...................................................................................................................21
5.1 Proppant Selection Methodology........................................................................................22
6 FracCADE Software .................................................................................................................28
6.1 The FGS Module ................................................................................................................29
6.2 The MLF Module ................................................................................................................29
6.3 The FracNPV Module .........................................................................................................30
6.4 The INVERSE Module........................................................................................................30
6.5 The PLACEMENT Module..................................................................................................31
6.5.1 PLACEMENT I ..........................................................................................................31
6.5.2 PLACEMENT II .........................................................................................................31
6.6 Additional FracCADE Modules ...........................................................................................31
7 Equations..................................................................................................................................33
7.1 Basic Equations..................................................................................................................33
7.2 Design Equations ...............................................................................................................34
7.3 Proppant Equations ............................................................................................................37
8 Proppant Flowback...................................................................................................................38
9 Refracturing ..............................................................................................................................39
9.1 Candidate Selection ...........................................................................................................40
9.2 Refracturing Design Methodology ......................................................................................41
9.2.1 Estimation of Recoverable Reserves ........................................................................41
9.2.2 Estimation of the Present Fracture Geometry...........................................................42
9.2.3 Determining the Average Reservoir Pressure...........................................................44
9.2.4 Estimating Production Response From Refracturing ................................................44
9.3 The Effects of Fractured Well Pressure Distributions on Design........................................44
9.4 Effects of Reservoir Pressure Changes .............................................................................48
9.5 Refracture Treatment Design Considerations ....................................................................50
9.5.1 Fracture Initiation ...................................................................................................... 50
9.5.2 Fracture Extension ....................................................................................................50
9.5.3 Fracture Containment ...............................................................................................52

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 7 of 33

9.5.4 Screenout ................................................................................................................. 52


9.5.5 Fracture-Fluid Recovery ........................................................................................... 52
Section 400 Figures
Fig. 1. Basic fracture treatment design. ....................................................................................... 4
Fig. 2. Ideal fracture treatment design. ........................................................................................ 5
Fig. 3. Logic and associated activities of an adaptive procedure for fracture treatment design.. 7
Fig. 4. The modelization concept relevant to treatment design. .................................................. 9
Fig. 5. Fracturing fluid selection guide for a gas well. ................................................................ 14
Fig. 6. Fracturing fluid selection guide for an oil well. ................................................................ 15
Fig. 7. Effect on productivity index ratio (Jo, unfractured) of 5-in. damage around fracture. ...... 17
Fig. 8. Friction pressure drop of various tubing and casing sizes for 30 lbm/1000 gal delayed
(dashed lines) and non-delayed (solid lines) borate fluids.............................................. 19
Fig. 9. Laminar and turbulent flow areas of viscoelastic fluids. .................................................. 19
Fig. 10. Fracture half-length requirements for a gas well........................................................... 23
Fig. 11. Retained permeability for linear fluids. .......................................................................... 24
Fig. 12. Retained permeability for borate-crosslinked fluids. ..................................................... 25
Fig. 13. Retained permeability for titanate-crosslinked fluids..................................................... 26
Fig. 14. Example of a production decline plot used to estimate recoverable reserves. ............. 42
Fig. 15. Example of a buildup data type-curve match. ............................................................... 43
Fig. 16. Example of a performance data type-curve match. ...................................................... 43
Fig. 17. Pressure contours around high-conductivity propped fractures.................................... 44
Fig. 18. The location of an infill well to improve recovery........................................................... 45
Fig. 19. Possible problem is fracture azimuth is unknown. ........................................................ 46
Fig. 20. Pressure contours around a short fracture or a fracture with low conductivity.............. 47
Fig. 21. Horizontal stress due to Poisson's ratio and pore pressure.......................................... 49
Fig. 22. Fracturing above a previous proppant-pack. ................................................................ 51
Section 400 Tables
Table 1. Core data example....................................................................................................... 12
Table 2. Rheology selection guidelines (pump time less than four hours)................................ 20
Table 3. Rheology selection guidelines (pump time greater than four hours)........................... 21
Table 4. Poppant Selection Guide ............................................................................................. 27
Section 500 Treatment Execution
1 Job Planning .............................................................................................................................. 4
1.1 Minimum Job Planning......................................................................................................... 4
1.2 Minimum Service Quality and Safety Standards.................................................................. 5
1.3 Operational Considerations ................................................................................................. 5
1.3.1 Mixing Fracturing Fluids ............................................................................................. 5
1.3.2 Equipment .................................................................................................................. 6
2 Safety......................................................................................................................................... 7
3 Environment............................................................................................................................... 7
4 Quality........................................................................................................................................ 7
4.1 Location Quality Assurance ................................................................................................. 7
4.2 Fracturing Fluid Kit............................................................................................................... 8

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 8 of 33

4.3 Testing References ..............................................................................................................8


5 Mixing and Metering of Additive Solutions..................................................................................9
5.1 Liquid Additives ....................................................................................................................9
5.1.1 Mixing Liquid-Additive Blends ...................................................................................11
5.2 Dry Additives in Solution.....................................................................................................11
5.2.1 Soluble Additives.......................................................................................................11
5.2.1.1 Mixing Solutions of Soluble Dry Additives ......................................................12
5.2.2 Insoluble Additives ....................................................................................................12
5.2.2.1 Mixing and Metering Fluid-Loss-Additive Slurries ..........................................13
5.3 Liquid-Additive Injection and Metering................................................................................13
5.3.1 Correcting the Additive Injection Rate to Proppant Concentration ............................14
5.3.2 Example Calculation for Mixing and Metering Additive Solutions .............................14
5.4 Water Solubilities of Various Dry Additives as a Function of Temperature ........................15
5.5 Methanol Solubilities of L10 as a Function of Temperature ...............................................17
5.6 Mixing and Metering Parameters for Various Solutions .....................................................17
6 Pressure Analysis During Fracturing Operations .....................................................................22
6.1 DSP Downhole Sensor Package (Real-Time Downhole Data Acquisition) ........................22
6.1.1 Fracturing Applications..............................................................................................23
6.1.2 System Components.................................................................................................23
6.1.2.1 Pressure and Temperature Gauges...............................................................24
6.1.2.2 Interface Module.............................................................................................25
6.2 Other Real-Time Data Acquisition Systems .......................................................................25
6.3 Downhole Memory Gauges ................................................................................................25
6.4 Additional Equations...........................................................................................................25
7 A Methodology For Improving Computed Bottomhole Pressures ............................................26
7.1 Recommendations..............................................................................................................28
7.2 Discussion ..........................................................................................................................28
7.2.1 Calibration Phase......................................................................................................29
7.2.1.1 Field Calibration .............................................................................................29
7.2.1.2 Office Calibration............................................................................................31
7.2.2 Application Phase .....................................................................................................32
7.3 Procedure A  Estimating the Closure Pressure and the Perforation/Near-Wellbore
Friction Pressure .....................................................................................33
7.4 Procedure B  Estimating the Fluid Friction Pressure and the Net Pressure ...................34
7.5 Procedure C  Estimating the Fluid Friction of Slurry Using Rate Changes .....................35
7.6 Procedure D  Estimating the Slurry Friction Pressure Using the PPR Software .............36
7.7 Real-Time Application of Procedure A, Procedure B and Procedure C .............................37
7.8 Field Example.....................................................................................................................37
7.8.1 Net Pressures From Uncalibrated Field Data ...........................................................37
7.8.1.1 Estimating Closure Pressure and Perforation Friction....................................40
7.8.1.2 Measuring Pipe Friction of Non-Proppant Laden Fluids.................................42
7.8.1.3 Measuring Pipe Friction of Slurries ................................................................47
7.8.1.4 Office Calibration  Job Playback with PPR Software ..................................50

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 9 of 33

7.8.1.5 PPR Hydrostatic Head Computation.............................................................. 60


8 Flowback Recommendations ................................................................................................... 63
8.1 Energized Fluids Flowback Procedure............................................................................... 63
8.2 Choke-Size Determination ................................................................................................. 64
8.2.1 Choke-Size Determination for Foam Flowback (Greater than 55-Quality) ............... 64
8.2.2 Choke-Size Determination for Energized (Nonfoamed) Fluids................................. 66
9 Contingency Plans ................................................................................................................... 69
9.1 Insufficient Pump Rate ....................................................................................................... 69
9.2 Proppant Delivery Failure .................................................................................................. 69
9.3 Equipment Malfunction ...................................................................................................... 69
9.4 Screenout........................................................................................................................... 69
Section 500 Figures
Fig. 1. Compatability of common Dowell fracturing additives..................................................... 10
Fig. 2. The DSP system. ............................................................................................................ 24
Fig. 3. Schematic of treatment with systematic rate changes. ................................................... 30
Fig. 4. Friction multiplier as function of sand concentration. ...................................................... 31
Fig. 5. Treatment data for example well  uncalibrated. .......................................................... 38
Fig. 6. Net pressure plot for example well  uncalibrated......................................................... 39
Fig. 7. Procedure A Treatment data. ...................................................................................... 41
Fig. 8. Friction pressure of fresh water, 5 1/2-in. casing. ........................................................... 42
Fig. 9. Procedure B  Treatment data, YF540HT..................................................................... 43
Fig. 10. Field calibrated fluid friction curve, YF540HT. .............................................................. 45
Fig. 11. Treatment data during rate change, WF110. ................................................................ 46
Fig. 12. Field calibrated fluid friction curve, WF110 . ................................................................. 47
Fig. 13. Treatment data during first shut-down with proppant.................................................... 48
Fig. 14. Treatment data during second shut-down with proppant.............................................. 49
Fig. 15. Treatment data from end of proppant to shut-down...................................................... 51
Fig. 16. Treatment data for example well  calibrated. ............................................................ 53
Fig. 17. Net pressure plot for example well  calibrated........................................................... 54
Fig. 18. Treatment data during rate change, WF110  calibrated............................................ 55
Fig. 19. Procedure B  Treatment data, YF540HT  calibrated.............................................. 56
Fig. 20. Treatment data during first shut-down with proppant  calibrated. ............................. 57
Fig. 21. Treatment data during second shut-down with proppant  calibrated......................... 58
Fig. 22. Treatment data from end of proppant to shut-down  calibrated. ............................... 59
Fig. 23. Comparison of calculated hydrostatic pressures. ......................................................... 61
Fig. 24. Treatment data for example well  improved hydrostatic pressure. ............................ 62
Fig. 25. Determining QZ............................................................................................................. 65
Section 500 Tables
Table 1. J218 ............................................................................................................................. 15
Table 2. J353 ............................................................................................................................. 15
Table 3. L10 ............................................................................................................................... 16
Table 4. M3 ................................................................................................................................ 16
Table 5. M117 ............................................................................................................................ 16
Table 6. L10 ............................................................................................................................... 17

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 10 of 33

Table 7. Aqueous J218 Solutions...............................................................................................17


Table 8. L10/Methenol Solutions ................................................................................................18
Table 9. Aqueous L10 Solutions.................................................................................................18
Table 10. Aqueous J353 Solutions.............................................................................................19
Table 11. Aqueous M3 Solutions................................................................................................19
Table 12. Aqueous M117 Solutions............................................................................................20
Table 13. J84 and J418 Slurries.................................................................................................20
Table 14. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with M2 ...................21
Table 15. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with U28..................21
Table 16. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with J465 or J474
Respectively...............................................................................................................21
Table 17. Aqueous Breaker Aid J466 Solution ...........................................................................22
Table 18. Summary of Procedures A through D.........................................................................28
Table 19. Pipe Friction as Function of Rate YF540HT, 5-1/2 in. Casing ....................................44
Table 20. Comparison of Actual and Estimated Pipe Friction Proppant Laden Fluid .................50
Table 21. Choke Coeffiecient versus Choke Size ......................................................................66
Table 22. Required Flowrate to Gas-Lift Water ..........................................................................67
Table 23. Flow of Water through Chokes ...................................................................................68
Section 600 Treatment Evaluation
1 Introductory Summary ................................................................................................................2
1.1 Treatment Evaluation Methodology......................................................................................2
1.2 Minimum Service Quality and Safety Standards ..................................................................3
2 Fracturing Pressure Analysis......................................................................................................3
2.1 Injection Pressure Interpretation...........................................................................................3
2.1.1 Nolte-Smith Plot ..........................................................................................................3
2.1.2 In-Situ Stress Requirements .......................................................................................7
2.1.3 References..................................................................................................................8
2.2 Pressure Decline Analysis ....................................................................................................8
2.2.1 References..................................................................................................................9
2.3 Fracture Height Prediction and Post-Treatment Measurements ..........................................9
2.3.1 Sonic Logs ..................................................................................................................9
2.3.2 References..................................................................................................................9
3 Treatment Performance Monitoring..........................................................................................10
3.1 Inverse Analysis of Treatment and Production Data Records ............................................10
3.1.1 Fracture Characterization Using the ZODIAC Software............................................11
3.1.1.1 Fracture Storage ............................................................................................12
3.1.1.2 Fracture Face Skin Damage ..........................................................................13
3.1.1.3 Variable Fracture Conductivity .......................................................................13
3.1.1.4 Reservoir Permeability Anisotropy .................................................................14
3.1.1.5 Finite Reservoirs ............................................................................................14
4 Production Evaluation...............................................................................................................22
4.1 References .........................................................................................................................23

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 11 of 33

Section 600 Figures


Fig. 1. Slope interpretation for the Nolte-Smith plot. .................................................................... 7
Fig. 2. Example of required stress contrasts................................................................................ 8
Fig. 3. High-conductivity fracture comparison. ............................................................................ 15
Fig. 4. Low-conductivity fracture comparison............................................................................. 15
Fig. 5. Fracture face skin damage, moderate-conductivity fracture. .......................................... 16
Fig. 6. Fracture face skin damage, finite-conductivity fracture................................................... 16
Fig. 7. Fracture face skin damage comparison, low-conducitivy fracture. ................................. 17
Fig. 8. Fracture skin damage comparison, low-conductivity fracture. ........................................ 17
Fig. 9. Finite-conductivity fracturecomparison, high average dimensionless conductivity. ........ 18
Fig. 10. Finite-conductivity comparison, moderate average fracture conductivity...................... 18
Fig. 11. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft.............. 19
Fig. 12. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft.............. 19
Fig. 13. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft..................... 20
Fig. 14. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft..................... 20
Fig. 15. ZODIAC software examples.......................................................................................... 21
Fig. 16. ZODIAC software examples.......................................................................................... 21
Fig. 17. ZODIAC software examples.......................................................................................... 22
Section 700 Techniques

Section 700.1 DataFRAC Service


1 Introductory Summary................................................................................................................ 6
1.1 Closure Test......................................................................................................................... 7
1.1.1 Closure Test in a Permeable Zone ............................................................................. 7
1.1.2 Closure Test in a Nonpermeable Zone....................................................................... 9
1.2 Calibration Test.................................................................................................................... 9
1.3 Applications........................................................................................................................ 10
2 Design...................................................................................................................................... 11
2.1 Preparatory Engineering .................................................................................................... 11
2.1.1 Breakdown/Diversion Treatment .............................................................................. 11
2.1.2 Preliminary Fracture Design ..................................................................................... 11
2.1.3 Fracture Height......................................................................................................... 11
2.1.4 Wellbore Logging...................................................................................................... 12
2.1.4.1 Temperature and Gamma-Ray Logs ............................................................. 12
2.1.4.2 Fracture-Height Logs ..................................................................................... 13
2.1.5 Perforating ................................................................................................................ 13
2.1.5.1 Wellbore Restrictions ..................................................................................... 13
2.1.5.2 Perforation Phasing ....................................................................................... 14
2.1.5.3 Perforation Size ............................................................................................. 14
2.2 Closure Test....................................................................................................................... 15
2.2.1 Fluid Selection .......................................................................................................... 15
2.2.2 Injection Rates and Number of Steps ....................................................................... 15
2.2.3 Step Duration............................................................................................................ 15
2.2.4 Flow-Back Rate ........................................................................................................ 16

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 12 of 33

2.3 Calibration Test ..................................................................................................................17


2.3.1 Fluid Selection...........................................................................................................17
2.3.1.1 Foam ..............................................................................................................17
2.3.2 Fluid Volume .............................................................................................................17
2.3.3 Fluid Break-Time.......................................................................................................18
2.3.4 Fluid-Loss Additives ..................................................................................................18
2.3.5 Duration of Pressure Decline ....................................................................................18
2.4 Special Considerations in the DataFRAC Design...............................................................18
2.4.1 The Influence of Wellbore Fluid ................................................................................18
2.4.2 Prepad.......................................................................................................................18
2.4.3 Closure Pressure less than Hydrostatic Pressure.....................................................19
2.4.4 Post-Job Wireline Surveys ........................................................................................19
2.5 Terminology........................................................................................................................19
2.5.1 Fracture Extension Pressure.....................................................................................19
2.5.2 Initial Shut-in Pressure ..............................................................................................19
2.5.3 Closure Pressure ......................................................................................................19
2.5.4 Rebound Pressure ....................................................................................................19
2.6 Equipment Requirements ...................................................................................................20
2.6.1 Monitoring Equipment ...............................................................................................20
2.6.2 Pumping Equipment..................................................................................................20
2.6.3 Pressure Measuring Equipment................................................................................20
2.6.3.1 Surface Measurement Methods .....................................................................20
2.6.3.2 Bottomhole Pressure Gauge Measurement ...................................................22
2.6.4 Treating Equipment...................................................................................................23
2.6.5 Flowback Equipment.................................................................................................23
2.6.5.1 Magnetic Flowmeters .....................................................................................23
2.6.5.2 Turbine Flowmeters........................................................................................23
2.6.5.3 Chokes and Gate Valves................................................................................23
3 Execution..................................................................................................................................24
3.1 Pre-Performance Guidelines ..............................................................................................24
3.2 Closure Test .......................................................................................................................27
3.2.1 Step-Rate Phase.......................................................................................................27
3.2.2 Flowback Phase........................................................................................................32
3.2.2.1 Flow Control ...................................................................................................32
3.2.2.2 Flowmeters.....................................................................................................34
3.2.3 Closure Test Modifications........................................................................................34
3.3 Calibration Test ..................................................................................................................35
3.3.1 Injection Phase..........................................................................................................35
3.3.2 Pressure-Decline Phase ...........................................................................................36
3.3.3 Contingency Plans ....................................................................................................36
4 Evaluation.................................................................................................................................36
4.1 Closure Test Analysis.........................................................................................................37
4.1.1 Step Rate  The BHP-Versus-Rate Plot..................................................................37
4.1.2 Flowback  The BHP-Versus-Time Plot ..................................................................37

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 13 of 33

4.1.3 Confirmation of Closure Pressure............................................................................. 38


4.1.4 Rebound Pressure.................................................................................................... 40
4.2 Calibration Injection for Fracture Geometry ....................................................................... 40
4.2.1 Elastic Fracture Compliance..................................................................................... 41
4.2.2 Pressure During Pumping......................................................................................... 43
4.2.2.1 Fluid Flow and Pressure in Fracture .............................................................. 43
4.2.2.2 Nolte-Smith Plot and Evolution of Pressure During Pumping ........................ 45
4.2.3 Deviations from Ideal Geometry ............................................................................... 46
4.2.3.1 Height Growth ................................................................................................ 46
4.2.3.2 Fissures ......................................................................................................... 47
4.2.3.3 T-Shape Fracture........................................................................................... 48
4.2.4 Pressure Capacity .................................................................................................... 49
4.2.5 Near-Wellbore Restriction......................................................................................... 50
4.2.6 Fracturing Pressure Interpretation Summary ........................................................... 53
4.2.6.1 Example of Radial Fracture ........................................................................... 54
4.2.6.2 Simulation of Pressure During Pumping and Decline .................................... 54
4.3 Calibration Decline for Fluid-Loss Behavior ....................................................................... 56
4.3.1 Review of Decline Analysis....................................................................................... 56
4.3.2 Volume Function g.................................................................................................... 58
4.3.3 Fluid Efficiency.......................................................................................................... 59
4.3.4 Decline Function G ................................................................................................... 61
4.3.5 Non-Ideal Behavior ................................................................................................... 64
4.3.5.1 Change in Fracture Penetration After Shut-in................................................ 64
4.3.5.2 Height Growth ................................................................................................ 65
4.3.5.3 Pressure-Dependent Leakoff ......................................................................... 66
4.3.5.4 Spurt .............................................................................................................. 69
4.3.5.5 Closure Pressure Change.............................................................................. 69
4.3.5.6 Compressible Fluids ...................................................................................... 71
4.3.6 Fluid Efficiency Based on Pressure Analysis............................................................ 72
4.3.7 Decline-Analysis Procedure...................................................................................... 73
4.3.8 Steps to Correct Decline Analysis Using the FracCADE Software ........................... 75
4.3.8.1 The DataFRAC Software ............................................................................... 76
4.3.8.2 G-plot Interpretation by the DataFRAC Software........................................... 77
4.3.8.3 Modulus, Height or Fracture Toughness Calibrations.................................... 77
4.3.8.4 The β Ratio .................................................................................................... 78
4.3.9 Post Proppant Fracture Analysis .............................................................................. 80
4.3.10 References ............................................................................................................. 81
Section 700.1 Figures
Fig. 1. The effect of proppant-pack damage and fracture length on fracture NPV. ..................... 6
Fig. 2. Fracture extension pressure (unequal time steps)............................................................ 7
Fig. 3. The typical closure test. .................................................................................................... 8
Fig. 4. The G-plot (idealized). .................................................................................................... 10
Fig. 5. Channel restriction at the wellbore.................................................................................. 13

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 14 of 33

Fig. 6. The relation of perforation diameter and proppant concentration. ..................................14


Fig. 7. The effects of differing flowback rates. ............................................................................16
Fig. 8. The change in surface pressure during closure in deep, hot wells..................................21
Fig. 9. Hydrostatic head changes during closure. ......................................................................22
Fig. 10. The DataFRAC Service rig-up when pumping conductive fluids. ..................................25
Fig. 11. The DataFRAC Service rig-up when pumping nonconductive fluids. ............................26
Fig. 12. Friction pressure of water in the tubing and casing. ......................................................28
Fig. 13. Friction pressure of water in the annulus.......................................................................29
Fig. 14. Friction pressure of brine in the tubing and casing........................................................29
Fig. 15. Friction pressure of brine in the annulus. ......................................................................30
Fig. 16. Friction pressure of diesel in the tubing and casing. .....................................................30
Fig. 17. Friction pressure of diesel in the annulus. .....................................................................31
Fig. 18. Flow rate versus differential pressure in perforations....................................................31
Fig. 19. Flowback test (after Nolte, 1982/1994)..........................................................................38
Fig. 20. Effect of closure on BHP versus square root of t and G- plots. .....................................39
Fig. 21. Rebound pressure; lower bound of closure pressure....................................................40
Fig. 22. Analogy of a pressurized crack to a pre-loaded spring. ................................................42
Fig. 23. Evolution of fracture geometry and pressure during pumping.......................................45
Fig. 24. Pressure and width for height growth through barriers (after Nolte, 1989)...................46
Fig. 25. Pressure and width for opening natural fissures (after Nolte, 1989). ...........................47
Fig. 26. Pressure and width for T-shape fracture (after Nolte, 1989). .......................................48
Fig. 27. Definition of pressure capacity from in-situ stresses. ....................................................50
Fig. 28. Stress state within the entrance of deviated well or stress............................................51
Fig. 29. Mohr circle of deviated well or stress. ...........................................................................52
Fig. 30. Nolte-Smith plot of fracturing pressure. .........................................................................53
Fig. 31. Net pressure with radial fracture (after Smith et al. 1987). ...........................................54
Fig. 32. Measured and simulated net pressure: opening natural fissures (after Nolte, 1982). ..55
Fig. 33. Example of fracturing-related pressures (after Nolte, 1982)..........................................56
Fig. 34. Schematic for fracture area and time. ...........................................................................57
Fig. 35. Dimensionless volume function for fracture closure (after Nolte, 1986). ......................59
Fig. 36. Efficiency from closure time for no proppant, no spurt loss during pumping and
other ideal assumptions given in Section 4.3.1 (after Nolte, 1986)...............................60
Fig. 37. Conceptual response of pressure decline versus Nolte time-function
(after Castillo, 1987). ....................................................................................................62
Fig. 38. Penetration change during shut-in (after Nolte, 1990)...................................................65
Fig. 39. Diagnostic for height growth from decline data (after Nolte, 1990). ..............................66
Fig. 40. Diagnostic for stress sensitive fissures from injection and decline (after Nolte, 1990). 67
Fig. 41. Decline analysis for filtrate and reservoir control leakoff (after Nolte, 1993). ...............68
Fig. 42. Stress change during injection/shut-in for Cc (after Nolte et. al., 1993). ........................70
Fig. 43. Relative volume change of gas (after Nolte et. al., 1993).............................................72
Fig. 44. Decline analysis using “¾” rule (after Nolte, 1990)........................................................74
Fig. 45. Pressure and flow rate in fracture before and after shut-in (after Nolte, 1986). ...........79
Fig. 46. Diagnostic for closing on proppant from decline data (after Nolte, 1990).....................80

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 15 of 33

Section 700.1 Tables


Table 1. Approximate Choke Settings For Flowback Of Oil-Base Fluids (Sg = 0.7) .................. 33
Table 2. Approximate Choke Settings for Flowback of Water-Base Fluids (Sg = 1.0)............... 34
Table 3. Interpolated Values of α Over the Full Range of n....................................................... 58
Table 4. Values of Decline Function "G" .................................................................................... 63
Table 5. Correction Factors f c As Function Of ∆tD ...................................................................... 75
Section 700.2 Foam Fracturing
1 Introductory Summary................................................................................................................ 2
1.1 Foam Properties .................................................................................................................. 3
1.2 Foam Types ......................................................................................................................... 3
1.3 Foam Stability ...................................................................................................................... 4
1.4 Applications.......................................................................................................................... 4
2 Design........................................................................................................................................ 4
2.1 Choosing a Foam................................................................................................................. 4
2.1.1 The Liquid Phase........................................................................................................ 5
2.1.1.1 Linear Polymers ............................................................................................... 5
2.1.1.2 Crosslinked Polymers ...................................................................................... 6
2.1.1.3 Hydrocarbons and Alcohols ............................................................................. 7
2.1.2 The Gas Phase........................................................................................................... 7
2.1.2.1 Gas Behavior ................................................................................................... 9
2.1.2.2 Gas Solubility ................................................................................................. 10
2.1.3 Foaming Agent Selection.......................................................................................... 11
2.1.3.1 Material Compatibility with Foaming Agents .................................................. 11
2.2 Foam Rheology.................................................................................................................. 12
2.3 Fluid-Loss Properties ......................................................................................................... 13
2.3.1 Two-Phase Behavior of the Foam ............................................................................ 13
2.3.2 Wall-Building Effects................................................................................................. 13
2.4 Conductivity Damage ......................................................................................................... 14
2.5 Foam Quality...................................................................................................................... 16
2.6 Foam Texture..................................................................................................................... 19
2.7 Proppant Compensation .................................................................................................... 19
2.7.1 No Proppant Compensation ..................................................................................... 20
2.7.2 Constant Bottomhole Quality .................................................................................... 21
2.7.3 Decreasing Bottomhole Quality ................................................................................ 21
2.7.4 Constant Internal Phase ........................................................................................... 22
2.8 Friction Pressure................................................................................................................ 22
2.9 Yield Stress ........................................................................................................................ 23
2.10 Limitations of Application ................................................................................................. 23
2.11 Job Design ....................................................................................................................... 23
2.12 Calculations ..................................................................................................................... 23
2.12.1 Pressures and Rates .............................................................................................. 23
2.12.2 Equipment Requirements ....................................................................................... 25
2.12.3 Material Requirements ........................................................................................... 26

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 16 of 33

3 Execution..................................................................................................................................26
3.1 Foam Generation ...............................................................................................................28
3.2 Material Balance.................................................................................................................30
Section 700.2 Figures
Fig. 1. The effect of polymer loading on foam viscosity (50% quality foam).................................5
Fig. 2. The effect of various polymers on foam stability. ..............................................................6
Fig. 3. The effect of foam quality on viscosity (StableFOAM fluid). ............................................12
Fig. 4. Leakoff of a foam into the rock matrix. ............................................................................13
Fig. 5. Dimensionless polymer concentration factor...................................................................15
Fig. 6. Polymer concentration versus proppant-pack retained permeability...............................16
Fig. 7. Bubble arrangements for various foam-quality ranges....................................................17
Fig. 8. Proppant concentration limits in foam fluids. ...................................................................18
Fig. 9. The effect of proppant compensation methods on bottomhole foam quality. ..................20
Fig. 10. Friction through perforations. ........................................................................................25
Fig. 11. Schematic of foam fracturing treatment.........................................................................27
Fig. 12. Laminar and turbulent flow areas of foamed fluids........................................................29
Fig. 13. Foam generator.............................................................................................................29
Section 700.2 Tables
Table 1. Summary of Foam Fracturing Fluids ..............................................................................3
Table 2. Comparison of Nitrogen and Carbon Dioxide.................................................................7
Table 3. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant ........................15
Section 700.3 RampGEL Service
1 Introductory Summary ................................................................................................................1
1.1 Application ............................................................................................................................2
2 Design ........................................................................................................................................3
2.1 Pad Fluid ..............................................................................................................................4
2.2 Fracturing Fluid Transporting Proppant................................................................................4
2.3 Computer-Aided Design .......................................................................................................4
3 EXECUTION...............................................................................................................................4
3.1 Batch Mixing .........................................................................................................................4
3.2 Continuous Mixing ................................................................................................................4
Section 700.3 Figures
Fig. 1. Polymer concentration design scheme for the RampGEL software. .................................5
Section 700.4 CleanFRAC Service
1 Introductory Summary ................................................................................................................2
1.1 Applications ..........................................................................................................................2
2 Design ........................................................................................................................................2
2.1 Well Candidate Selection .....................................................................................................3
2.2 EB-Clean J475 Breaker........................................................................................................3
2.2.1 Mechanism of Breaker Release ..................................................................................5
2.3 Breaker Selection .................................................................................................................6
2.4 Job Design............................................................................................................................7

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 17 of 33

2.4.1 CleanFRAC Service Design Methods......................................................................... 7


2.4.2 Summary of Software Features for the CleanFRAC Service Design.......................... 7
2.4.3 Design Procedure Using a Manual Calculation .......................................................... 8
2.4.4 Design Procedure Using the FracCADE Software Exclusively or in Combination
With a Manual Calculation........................................................................................ 11
3 Execution ................................................................................................................................. 12
3.1 Metering the J475 Using the Auger-Type, Dry-Additive Feeder ........................................ 12
3.1.1 Procedure for Scheduling the J475 Using the Auger-Type Dry-Additive Feeder ..... 13
3.2 Metering the J218 as an Aqueous Solution ....................................................................... 14
3.2.1 Procedure for Calculating Mixing and Metering Parameters for J218 Solutions ...... 14
4 Examples ................................................................................................................................. 16
4.1 Example Design Using a Manual Calculation .................................................................... 16
4.2 Example Procedure for Scheduling the J475 Using the Auger-Type
Dry-Additive Feeder........................................................................................................... 18
4.3 Example Procedure for Calculating Mixing and Metering Parameters for
J218 Solutions ................................................................................................................... 19

Section 700.4 Figures


Fig. 1. Effect of breaker concentration on retained viscosity for a borate-crosslinked guar
fluid at 160°F (71°C).......................................................................................................... 4
Fig. 2. Effect of breaker concentration on retained proppant-pack permeability.......................... 5
Fig. 3. Effects of proppant concentration and porosity on postclosure polymer concentration. ... 9
Fig. 4. Influence of ammonium persulfate breaker on retained proppant-pack permeability
for linear fracturing fluids................................................................................................... 9
Fig. 5. Influence of ammonium persulfate breaker on retained proppant-pack permeability
for borate-crosslinked fracturing fluids. ........................................................................... 10
Fig. 6. Influence of ammonium persulfate breaker on retained proppant-pack permeability
for organometallic-crosslinked fracturing fluids............................................................... 10
Section 700.4 Tables
Table 1. J475 Release Levels (no closure stress) ....................................................................... 6
Table 2. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant.......................... 8
Table 3. J475 Metering Rates Auger-Type, Dry-Additive Feeder .............................................. 12
Section 700.5 Fracture-Height-Containment Services
1 Introductory Summary................................................................................................................ 2
1.1 Fracture Height Prediction ................................................................................................... 3
1.2 Fracture Penetration ............................................................................................................ 3
1.3 Fracture Evolution................................................................................................................ 4
1.4 Fracture Height-Growth and Containment ........................................................................... 7
1.5 Fracture-Height-Containment Services................................................................................ 9
2 DIVERTAFRAC Service........................................................................................................... 10
2.1 Discussion.......................................................................................................................... 10
2.2 Design Methodology .......................................................................................................... 11

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 18 of 33

2.2.1 Design Example ........................................................................................................13


2.3 Execution Methodology ......................................................................................................15
2.4 DIVERTAFRAC Fluid Using Water Control Agent S41 as a Diverting Material..................15
2.4.1 Spacers .....................................................................................................................15
2.4.2 Dilution with Sand .....................................................................................................15
2.4.3 Additional Information ...............................................................................................15
3 INVERTAFRAC Service ...........................................................................................................16
3.1 Design Methodology...........................................................................................................16
3.2 Execution Methodology ......................................................................................................17
3.2.1 Field Mixing Procedures............................................................................................18
3.2.2 Additional Mixing Techniques ...................................................................................19
4 Computer-Aided Design ...........................................................................................................20
4.1 Additional Computer-Aided Job Design Information...........................................................22
Section 700.5 Figures
Fig. 1. First phase of evolution. ....................................................................................................4
Fig. 2. Illustration of fracture growth. ............................................................................................5
Fig. 3. Linear plot of pressure.......................................................................................................6
Fig. 4. Log-Log plot of net pressure. ............................................................................................6
Fig. 5. Pressure and width for growth through barriers. ...............................................................7
Fig. 6. Height control and conventional treatment for an offset well.............................................8
Fig. 7. Principles of fracture-height-containment techniques........................................................9
Fig. 8. J423 rise rate versus fluid viscosity (static conditions). ...................................................17
Fig. 9. J423 flow rate versus area. .............................................................................................19
Section 700.5 Tables
Table 1. J423 Addition Rate .......................................................................................................18
Section 700.6 Breakdown Techniques
1 Introductory Summary ................................................................................................................2
2 Applications ................................................................................................................................3
3 Treatment Design .......................................................................................................................3
3.1 Fracture Gradient .................................................................................................................5
3.2 Fluid Selection ......................................................................................................................5
3.2.1 Solvent Selection Guidelines.......................................................................................6
3.3 Types of Formation Damage ................................................................................................8
3.4 Fluid-Loss Control ..............................................................................................................13
3.5 Fluid Volumes.....................................................................................................................13
3.6 Fluid Diversion....................................................................................................................14
3.6.1 Ball Sealers ...............................................................................................................14
3.6.2 Chemical Diverting Agents ........................................................................................15
4 Execution..................................................................................................................................16
5 Evaluation.................................................................................................................................17
6 Examples..................................................................................................................................18
6.1 Example No. 1 ....................................................................................................................18

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 19 of 33

6.2 Example No. 2 ................................................................................................................... 20


Section 700.6 Figures
Fig. 1. Ball catcher. .................................................................................................................... 17
Fig. 2. Friction loss of 15% HCl in 2.875 in. tubing. ................................................................... 22
Section 700.6 Tables
Table 1. Breakdown Design Methodology ................................................................................... 4
Table 2. The Dowell Solvent Formulations Most Commonly Used for Breakdown Fluids (The
Stimulation Materials  Acidizing Manual provides information for these systems)...... 7
Table 3. Summary of Formation Damage.................................................................................... 9
Section 700.7 Diverting Techniques
1 Introductory Summary................................................................................................................ 2
2 Application ................................................................................................................................. 4
2.1 Ball Sealers .......................................................................................................................... 5
2.1.1 Buoyant Ball Sealers .................................................................................................. 5
2.1.2 Conventional Ball Sealers........................................................................................... 8
2.1.3 Temperature and Pressure Effects........................................................................... 13
2.1.4 Effects of Various Fluids ........................................................................................... 13
2.2 Limited Entry ...................................................................................................................... 14
2.2.1 Perforation Erosion ................................................................................................... 18
2.2.2 Contributing Factors ................................................................................................. 19
2.2.2.1 Casing Size.................................................................................................... 19
2.2.2.2 Number and Size of Perforations................................................................... 19
2.2.2.3 Differences in Fracturing Pressures............................................................... 21
2.2.2.4 Pressure Limitations ...................................................................................... 21
2.2.2.5 Hydraulic Horsepower Requirements ............................................................ 21
2.2.2.6 Perforation Friction Pressure ......................................................................... 21
2.2.2.7 Breakdown Technique ................................................................................... 22
2.2.2.8 Determination of Formation Fracturing Pressure........................................... 22
2.2.3 Design ...................................................................................................................... 22
2.3 Single-Point Entry .............................................................................................................. 33
2.4 Plugback (Pine Island) Technique ..................................................................................... 34
2.5 Slurried Solids .................................................................................................................... 36
2.6 Viscous Fluids .................................................................................................................... 37
2.7 Baffles and Balls ................................................................................................................ 38
2.8 Bridge Plugs....................................................................................................................... 40
Section 700.7 Figures
Fig. 1. Rising velocity of 7/8-in. OD buoyant ball sealers............................................................. 7
Fig. 2. This graph provides information necessary for calculating the injection rate to
“balance” buoyant balls in different sizes of pipe. ............................................................. 7
Fig. 3. Example printout of the PERFBALL ball-sealer-performance computer software. ........... 9
Fig. 4. This graph contrasts the pump rate versus fluid velocity in different sizes of pipes........ 12

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 20 of 33

Fig. 5. This graph contrasts the pump rate versus fluid velocity in different sizes of
perforations. ....................................................................................................................12
Fig. 6. Core selection guide for temperature and pressure. .......................................................13
Fig. 7. Injection rate increases thus friction pressure increases.................................................15
Fig. 8. Nomograph for determining friction through a perforation...............................................16
Fig. 9. Reference to determine the injection rate per perforation using fresh water...................17
Fig. 10. Higher density fluids exhibit higher perforation friction pressures. ................................18
Fig. 11. Change in perforation friction versus perforation coefficient. ........................................20
Fig. 12. Gamma ray-sonic log. ...................................................................................................23
Fig. 13. Friction of WF140 in 4 1/2 and 5 1/2 in. pipe.................................................................25
Fig. 14. Plugback Pine Island technique. ...................................................................................34
Fig. 15. Two-stage fracturing procedure, using baffle and ball technique. .................................39
Section 700.7 Tables
Table 1. Perforation Size versus Ball-Sealer Core Size ...............................................................5
Table 2. Example for Selecting a Perforation Schedule for Multiple Zones Having Different
Bottomhole Fracturing Pressures .................................................................................32
Table 3. Diverting Agents ...........................................................................................................36
Section 700.8 CleanFLOW Technology
1 Introductory Summary ................................................................................................................1
2 The CleanFLOW System............................................................................................................2
2.1 Optimized Breaker Composition...........................................................................................2
2.2 J495 CleanFLOW Additive ...................................................................................................3
2.3 CleanFLOW Application .......................................................................................................4
2.3.1 Temperature ...............................................................................................................4
2.3.2 J495 Concentrations ...................................................................................................4
2.3.3 Breaker Interactions ....................................................................................................4
2.3.4 Material Handling ........................................................................................................5
2.3.5 Field Results ...............................................................................................................5
3 Conductivity Measurements .......................................................................................................5
3.1 Polymer Concentration .........................................................................................................6
3.2 Pressure-Drop Measurements .............................................................................................7
3.3 Other Parameters .................................................................................................................7
3.4 Summary ..............................................................................................................................7
Section 700.8 Tables
Table 1. Breaker Schedule Design Example ................................................................................3
Section 700.9 PropNET Technology
1 Introductory Summary ................................................................................................................2
2 Design ........................................................................................................................................6
2.1 Flowback Stability.................................................................................................................6
2.2 Two-Phase Flow.................................................................................................................10
2.3 Effective Proppant-Pack Stress Cycling .............................................................................11
2.4 Effect of Fluid Viscosity on Proppant-Pack Stability ...........................................................12
2.5 Proppant-Pack Permeability ...............................................................................................13

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 21 of 33

2.6 PropNET Lifetime............................................................................................................... 19


2.7 Stability in Acids ................................................................................................................. 21
2.8 Effect on Proppant Settling ................................................................................................ 21
2.9 Fiber Breakage During Treatments.................................................................................... 23
2.10 Case Histories.................................................................................................................. 25
3 PropNET References ............................................................................................................... 29
4 Appendix A - Flow Test Apparatus........................................................................................... 30
4.1 Fracture Geometry............................................................................................................. 30
4.2 Perforation Geometry......................................................................................................... 30
4.3 Tube Geometry .................................................................................................................. 30
Section 700.9 Figures
Fig. 1. A molehole formed in a proppant pack of 20/40-mesh sand (FracTech test),
viewed from the 5.25 in. by 5.25 in. face. ....................................................................... 10
Fig. 2. Pressure at failure versus nitrogen flow concentration. .................................................. 11
Fig. 3. Effect of fluid viscosity on pack stability. ......................................................................... 13
Fig. 4. The effect of closure stress on proppant-pack permeability with and without J500. ....... 18
Fig. 5. The effect of J501 concentration on the permeability of 16/20 Carbolite.
(Fluid: corn syrup.) .......................................................................................................... 19
Fig. 6. Predicted proppant-pack strength with aging time for J500. ........................................... 20
Fig. 7a. Proppant settling in WF150 fluid. .................................................................................. 22
Fig. 7b. Proppant settling in WF150 fluid.
Fig. 8. The effect of J501 on proppant settling in WF175 fluid under dynamic conditions. ........ 23
Fig. 9. Resin-coated proppant containing PropNET fibers after pumping and removal
from a well....................................................................................................................... 24
Fig. 10. PropNET fibers after pumping, removal from a well and cleaning to remove oil. ........ 25
Fig. 11. South Texas well, 15% J501 tail-in. .............................................................................. 26
Fig. 12. Fluid returns, South Texas offset wells. ........................................................................ 27
Fig. 13. Polymer returns, offset wells. ........................................................................................ 28
Fig. 14. Fracture geometry test apparatus................................................................................. 31
Fig. 15. Perforation geometry test apparatus............................................................................. 32
Fig. 16. Tube geometry test apparatus. ..................................................................................... 32
Section 700.9 Tables
Table 1. Summary of Flowback and Production Rates for the Highest-Rate PropNET
Wells to date................................................................................................................... 5
Table 2. Resistance to Flowback in the Fracture Geometry ........................................................ 7
Table 3. Resistance to Flowback in the Perforation Geometry .................................................... 8
Table 4. Failure of Proppant Packs Containing J501 in Cyclic Loading..................................... 12
Table 5. Permeabilities of Various ISP and Curable RCP Proppant Packs ............................... 14
Table 6. Permeabilities of Various Sand and Curable RCP Packs ............................................ 14
Table 7. Permeabilities of Precured RCP Packs........................................................................ 15
Table 8. Permeabilities of 20/40-Mesh Ottawa Sand Packs ...................................................... 15
Table 9. Permeabilities of 12/20-Mesh Ottawa Sand Packs ...................................................... 16
Table 10. Permeabilities of 16/20-Mesh Carbolite Packs .......................................................... 17
Table 11. Permeabilities of 20/40-Mesh Interprop Plus Packs .................................................. 17

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 22 of 33

Section 700.9 Supplement


1 Wellbore Cleanup .......................................................................................................................1
2 Flowback, Swabbing...................................................................................................................1
3 Shut-in, Clean-up, and Production Rates ...................................................................................2
4 Appendix.....................................................................................................................................4
Section 700.9 Supplement Figures
Fig. 1. Maximum swab-cable speed vs casing ID and number of perforations. ...........................4
Section 700.9 Supplement Tables
Table 1. Maximum Worldwide PropNET Flowback and Production Rates...................................5
Section 800 HyPerSTIM Service
1 Introduction.................................................................................................................................4
1.1 Objectives.............................................................................................................................4
1.2 Applications ..........................................................................................................................5
1.3 Limitations of Application......................................................................................................6
2 Design ........................................................................................................................................6
2.1 Candidate Selection .............................................................................................................7
2.2 Characterization of Formation Mechanical Properties..........................................................9
2.3 Design Basis.......................................................................................................................10
2.4 Fluid Selection ....................................................................................................................15
2.4.1 Fluid-Loss Control .....................................................................................................16
2.4.1.1 Pressure Effects .............................................................................................18
2.4.1.2 Temperature Effects.......................................................................................19
2.4.1.3 Effects of Fluid Viscosity and Polymer ...........................................................20
2.4.1.4 Effects of Fluid-Loss Additives .......................................................................21
2.4.1.5 Fluid Selection and Fluid-Loss Control...........................................................21
2.4.2 The DataFRAC Service Application ..........................................................................21
2.5 Proppant Selection and Fracture Conductivity ...................................................................22
2.5.1 Embedment...............................................................................................................23
2.5.1.1 Spalling...........................................................................................................25
2.5.1.2 Impact on Permeability...................................................................................26
2.5.2 Non-Darcy Flow ........................................................................................................27
2.5.2.1 Determination of the Inertial Flow Coefficient.................................................28
2.5.2.2 Non-Darcy Flow Correction of Dimensionless Fracture Conductivity.............32
2.5.2.3 Proppant Selection Using Manual Calculation ...............................................33
2.5.2.4 Computer-Aided Proppant Selection ..............................................................35
2.5.2.5 Proppant Selection Using the FracCADE Software........................................37
2.5.2.6 Proppant Selection Summary.........................................................................37
2.5.3 Formation Sand and Fines........................................................................................37
2.5.3.1 Control of Formation Fines and Sand ............................................................40
2.5.4 Proppant Flowback Control.......................................................................................40
2.6 FracCADE Software ...........................................................................................................44
2.6.1 FracNPV and QUICK Modules..................................................................................44
2.6.2 The FORECAST Module...........................................................................................47

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 23 of 33

2.6.3 The PLACEMENT II Simulator ................................................................................. 49


3 Execution ................................................................................................................................. 54
3.1 Batch-Mix Operations ........................................................................................................ 54
3.2 Continuous-Mix Operations ............................................................................................... 55
4 Evaluation ................................................................................................................................ 55
4.1 Prats’ Correlation ............................................................................................................... 56
4.2 Modified McGuire-Sikora Correlation ................................................................................. 56
5 Fluid-Loss Data........................................................................................................................ 64
5.1 WF120 (J164) Containing 25 lbm J478/1000 gal and 25 lbm J418/1000 gal
BHST=150°F (66°C), Pressure=1000 psi .......................................................................... 64
5.2 WF160 (J164) Without Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 65
5.3 WF160 (J164) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 66
5.4 WF110 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure- 1000 psi ............................................................................................................ 67
5.5 WF120 (J424) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 68
5.6 WF130 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 69
5.7 WF140 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 70
5.8 WF160 (J424) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ............................................................................................................. 71
5.9 YF140 (J424)  BHST=150°F (66°C), Pressure-1000 psi................................................ 72
5.10 YF140 (J424)  BHST=150°F (66°C), Pressure-1000 psi.............................................. 73
5.11 YF140 (J424)  BHST=175°F (79°C), Pressure-1000 psi.............................................. 74
5.12 YF140HTD (J424)  BHST=200°F (93°C), Pressure-1000 psi ...................................... 75
5.13 YF140HTD (J424)  BHST=250°F (121°C), Pressure-1000 psi .................................... 76
Section 800 Figures
Fig. 1. Production rate sensitivity to skin...................................................................................... 8
Fig. 2. IPR curve sensitivity to skin. ............................................................................................. 8
Fig. 3. The effect on shifting an 80% damage collar. ................................................................. 11
Fig. 4. Productivity-increase curves. .......................................................................................... 12
Fig. 5. Effective wellbore radius for pseudo-radial flow.............................................................. 14
Fig. 6. Fluid-loss data for YF140................................................................................................ 17
Fig. 7. Pressure gradient through a sand pack versus gas flow rate, darcy and non-darcy flow.27
Fig. 8. Total pressure drawdown versus transit time, sanding prediction. ................................. 39
Fig. 9. Curable resin-coated proppant compressive strength required to prevent flowback. .... 44
Fig. 10. Proppant Editor. ............................................................................................................ 45
Fig. 11. FracNPV Input. ............................................................................................................. 45
Fig. 12. Equivalent wellbore radius and pseudo-skin................................................................. 47
Fig. 13. PRODUCTION FORECAST input................................................................................. 47
Fig. 14. Production simulation, non-darcy flow. ......................................................................... 49

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 24 of 33

Fig. 15. ROCK input. ..................................................................................................................50


Fig. 16. ZONES — layer data input. ...........................................................................................51
Fig. 17. PLACEMENT SIMULATOR — conventional design, 20/40-mesh sand,
1400 gal pad. ................................................................................................................57
Fig. 18. PLACEMENT OUTPUT — conventional design, 20/40-mesh sand. .............................57
Fig. 19. PLACEMENT SIMULATOR — P3D tip-screenout design, 20/40-mesh sand,
1600 gal pad. ................................................................................................................58
Fig. 20. PLACEMENT OUTPUT — P3D tip-screenout design, 20/40-mesh sand. ....................58
Fig. 21. PLACEMENT SIMULATOR — conventional design, 12/20-mesh sand, 1800 gal pad.59
Fig. 22. PLACEMENT OUTPUT — conventional design, 12/20- mesh sand. ............................59
Fig. 23. PLACEMENT SIMULATOR — P3D tip-screenout design, 12/20-mesh sand,
3500 gal pad. ................................................................................................................60
Fig. 24. PLACEMENT OUTPUT — P3D tip-screenout design, 12/20-mesh sand. ...................60
Fig. 25. Stage front propogation. ................................................................................................61
Fig. 26. Fracture height profile ...................................................................................................61
Fig. 27. Wellbore fracture width profile. ......................................................................................62
Fig. 28. Fracture height growth history. ...................................................................................... 62
Fig. 29. Fracturing (net) pressure profile. ................................................................................... 63
Section 800 Tables
Table 1. Water Viscosity at Temperature ...................................................................................20
Table 2. Dry Proppant Pack Intertial Coefficient Factors............................................................31
Table 3. Proppant Selection With Embedment and Non-Darcy Flow .........................................35
Section 900 Acid Fracturing
1 Principles of Acid Fracturing.......................................................................................................2
1.1 Fracture Length and Fracture Conductivity ..........................................................................4
1.2 Factors Affecting Acid Behavior in Carbonate Reservoirs....................................................6
1.2.1 Acid Type, Strength, and Volume ...............................................................................6
1.2.2 Acid Leakoff ................................................................................................................9
1.2.3 Controlling Acid Leakoff ............................................................................................14
1.2.4 Acid Reaction Rate ...................................................................................................15
1.2.5 Acid Spending Time ..................................................................................................17
2 Treatment Design Fundamentals for Acid Fracturing...............................................................18
2.1 Achieving Acid Penetration.................................................................................................19
2.2 When Acid Fracture Length Should be Maximum ..............................................................19
2.3 When Acid Fracture Length Should be Limited ..................................................................19
2.4 Maximizing The Injection Rate ...........................................................................................20

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 25 of 33

2.5 Optimizing Conductivity and Etched Fracture Length ........................................................ 20


2.6 Effective Acid Concentration .............................................................................................. 21
2.7 Selecting Fluids for Deeper Acid Penetration .................................................................... 22
2.8 Determination of Leakoff Coefficients ................................................................................ 24
2.8.1 Methodology ............................................................................................................. 24
2.8.2 Example Calculation ................................................................................................. 29
2.8.3 Notes ........................................................................................................................ 31
2.9 Cooldown ........................................................................................................................... 31
2.10 Retarded Acid .................................................................................................................. 33
2.11 Viscous Fingering ............................................................................................................ 36
2.12 Summary of Treatment Design Fundamentals for Acid Fracturing .................................. 37
Section 900 Figures
Fig. 1. Conductivity ratio versus increase in folds. ....................................................................... 5
Fig. 2. Acid spending in carbonate rock. ...................................................................................... 8
Fig. 3. Casting of a typical wormhole pattern in limestone......................................................... 12
Fig. 4. Test results, carbonate cores and wormholes. ............................................................... 12
Fig. 5. Test results, carbonate cores and wormholes. ............................................................... 13
Fig. 6. Test results, carbonate cores and wormholes. ............................................................... 13
Fig. 7. Viscosity of fresh water versus temperature. .................................................................. 25
Fig. 8. Spending time versus HCl concentration at static conditions. ........................................ 35
Fig. 9. Temperature versus retardation factor at dynamic test conditions. ................................ 36
Section 900 Tables
Table 1. Factors that Influence Fracture Conductivity and Fracture Penetration when Acid
Fracturing a Carbonate Reservoir .................................................................................. 6
Table 2. Acid Types and Strengths Common to Oilfield Operations............................................ 9
Table 3. Total Minimum Leakoff Coefficients versus Temperature and Permeability ................ 26
Table 4. Scale Factor Base Values (SFb)................................................................................... 27
Table 5. Scale Factor Corrections ............................................................................................. 28
Table 6. Retardation Factor Selection Guidelines...................................................................... 34
Section 1000 Horizontal Wells
1 Introductory Summary................................................................................................................ 4
1.1 Candidate Reservoirs for Horizontal Wells .......................................................................... 4
1.1.1 Naturally-Fractured Reservoirs................................................................................... 4
1.1.2 Matrix-Permeability Reservoirs................................................................................... 4
1.1.2.1 Vertical Permeability ........................................................................................ 5
1.1.2.2 Skin Damage ................................................................................................... 6
1.2 Fracture Performance and Wellbore Orientation ................................................................. 7
2 Fracture Orientation ................................................................................................................... 8
3 Fractured Horizontal Well Performance ................................................................................... 10
3.1 Longitudinal Fracture ......................................................................................................... 10
3.2 Orthogonal Fractures ......................................................................................................... 11
3.3 Choke Skin Effect............................................................................................................... 13
3.3.1 Fracture Reorientation — Choke Effect.................................................................... 14

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 26 of 33

3.4 Coning Effects ....................................................................................................................15


3.4.1 Comparison of Fractured and Nonfractured Reservoir .............................................15
3.4.2 Effect of the Distance From the Fracture to the Water Zone ....................................15
4 Rock Mechanical Properties.....................................................................................................17
4.1 Openhole Wellbore Stability ...............................................................................................17
4.2 Shear Failure ......................................................................................................................18
4.3 Tensile Failure ....................................................................................................................18
4.4 Matrix Collapse...................................................................................................................18
4.5 Cased-Hole Wellbore Stability ............................................................................................19
4.6 Stress and Deformation Analysis .......................................................................................20
5 Fracture Initiation And Propagation ..........................................................................................20
5.1 Initiation Pressure...............................................................................................................20
5.2 Fracture Initiation................................................................................................................21
5.3 Fracture Propagation..........................................................................................................23
5.4 Longitudinal Fractures ........................................................................................................23
5.5 Angled Fractures ................................................................................................................24
5.6 Transverse Fractures .........................................................................................................25
5.7 Controlling Fracture Reorientation......................................................................................26
6 Perforating ................................................................................................................................27
7 Treatment Design .....................................................................................................................29
7.1 Net Present Value Analysis ................................................................................................29
7.1.1 Calculating the NPV of Orthogonal Fractures ...........................................................29
7.1.2 Horizontal Well Production Prediction .......................................................................31
7.2 Fracture Height...................................................................................................................34
7.3 Fracture Orientation ...........................................................................................................34
7.4 Fracture Length and Conductivity.......................................................................................34
7.5 Pump Rate..........................................................................................................................34
7.6 Fracturing Fluid Selection...................................................................................................34
7.7 Proppant Selection .............................................................................................................35
7.7.1 Mesh Range..............................................................................................................35
7.7.2 Proppant Type...........................................................................................................35
7.7.3 Proppant Concentration ............................................................................................35
8 Execution..................................................................................................................................35
8.1 Perforating ..........................................................................................................................35
8.1.1 When to Perforate .....................................................................................................36
8.2 Wellbore Isolation Between Fractures................................................................................36
8.2.1 Isolation Using Mechanical Tools..............................................................................37
8.2.2 Isolation Using Proppant Plugs .................................................................................37
8.2.2.1 Intentional Screenout .....................................................................................37
8.2.2.2 Multidensity/Multimesh Proppant ...................................................................38
8.2.3 Isolation Using Viscous Plugs ...................................................................................38
8.2.4 Wellbore Isolation in an Openhole Completion .........................................................39
8.3 Flowback ............................................................................................................................39
9 Overview of the Horizontal Well Treatment-Design Procedure ................................................40
10 Case History ...........................................................................................................................41

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 27 of 33

Section 1000 Figures


Fig. 1. Productivity index ratios for horizontal versus vertical wells. ............................................ 6
Fig. 2. A longitudinal fracture. ...................................................................................................... 8
Fig. 3. An orthogonal fracture. ..................................................................................................... 9
Fig. 4. A fracture propagating at an angle to the wellbore. .......................................................... 9
Fig. 5. Productivity index ratios of vertical well/vertical fracture and horizontal well with a
longitudinal fracture......................................................................................................... 10
Fig. 6. NPV analysis for a vertical well. ...................................................................................... 12
Fig. 7. NPV analysis for one orthogonal fracture. ...................................................................... 12
Fig. 8. Water breakthrough time versus the position of the fracture with respect to
the oil/water contact. ....................................................................................................... 16
Fig. 9. Production rate versus water-cut. ................................................................................... 16
Fig. 10. Fracture conductivity versus water-cut. ........................................................................ 17
Fig. 11. Mohr failure envelope for matrix collapse. .................................................................... 18
Fig. 12. Stress distribution around a perforation. ....................................................................... 20
Fig. 13. Initiation pressure as a function of α and the borehole inclination. ............................... 21
Fig. 14. Horizontal well configuration in the in-situ stress field................................................... 21
Fig. 15. Initiation points and fracture orientation on the borehole. ............................................. 22
Fig. 16. Fracture initiation pressure. .......................................................................................... 22
Fig. 17. The effect of distance between collinear fractures on maximum fracture width............ 24
Fig. 18. Fracture rotation angle versus spacing......................................................................... 25
Fig. 19. Width and excess pressure as a function of spacing for parallel, transverse
and radial fractures........................................................................................................ 26
Fig. 20. Radius of fracture reorientation as a function of the ratio between the maximum
and minimum horizontal stresses. ................................................................................. 27
Fig. 21. Critical distance between perforation versus well orientation. ...................................... 28
Fig. 22. Single phase flow .......................................................................................................... 33
Fig. 23. NPV analysis of the number of orthogonal fractures..................................................... 42
Fig. 24. Actual versus predicted fluid production. ...................................................................... 43
Section 1000 Tables
Table 1. Performance Comparison Of Vertical And Horizontal Wells With Fractures................ 13
Table 2. Pseudoskin Factor Correlation Contstants................................................................... 32
Table 3. Overview of Fracture Treatment Design Considerations for Horizontal Wells ............. 40
Table 4. Stress Measurement Techniques ................................................................................ 41
Section 1100 Appendix A - Fracturing Design-Execution-Evaluation Example
1 Well Data ................................................................................................................................... 3
2 Reservoir Evaluation .................................................................................................................. 3
3 DataFRAC Service................................................................................................................... 11
4 Treatment Design .................................................................................................................... 22
4.1 Fracturing Fluid Selection .................................................................................................. 22
4.2 Proppant Selection............................................................................................................. 22
4.3 Fracture Length Optimization............................................................................................. 24
4.4 In-Situ Stress Data ............................................................................................................. 27
4.5 Approximate Pumping Schedule........................................................................................ 28

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 28 of 33

4.6 Placement Design ..............................................................................................................32


4.7 Production Forecast ...........................................................................................................35
5 Treatment Execution ................................................................................................................40
6 Treatment Evaluation ...............................................................................................................42
7 Post-Fracture Well-Test Analysis .............................................................................................44
Section 1100 Figures
Fig. 1. Review output for all transients. ........................................................................................5
Fig. 2. Diagnostic plot - Transient 1..............................................................................................7
Fig. 3. Wellbore storage - Transient 1. .........................................................................................8
Fig. 4. Horner plot - Transient 1. ..................................................................................................9
Fig. 5. Multi-rate type curve and derivative match......................................................................10
Fig. 6. Verification plot. ...............................................................................................................10
Fig. 7. Production decline. ..........................................................................................................11
Fig. 8. NODAL analysis. .............................................................................................................12
Fig. 9. DataFRAC job record. .....................................................................................................13
Fig. 10. Step-rate test. ................................................................................................................13
Fig. 11. DataFRAC job record (replotted)...................................................................................14
Fig. 12. Pressure analysis ..........................................................................................................14
Fig. 13. Closure pressure estimation.......................................................................................... 16
Fig. 14. “G” plot. .........................................................................................................................18
Fig. 15. Net Present Value. ........................................................................................................24
Fig. 16. Cumulative production...................................................................................................25
Fig. 17. Fracture height-growth history....................................................................................... 35
Fig. 18. Wellbore fracture width profile. ......................................................................................35
Fig. 19. Fracture height profile. ..................................................................................................36
Fig. 20. Production decline. ........................................................................................................40
Fig. 21. Cumulative production...................................................................................................40
Fig. 22. Post-fracture IPR curves (NODAL analysis)..................................................................41
Fig. 23. Job record. ....................................................................................................................42
Fig. 24. Pressure analysis (Nolte-Smith plot). ............................................................................43
Fig. 25. Job record. ....................................................................................................................44
Fig. 26. Pressure analysis. .........................................................................................................44
Fig. 27. Linear flow regime. ........................................................................................................47
Fig. 28. Vertical fracture, linear flow - Transient 1. .....................................................................48
Fig. 29. Multi-rate type curve and derivative match....................................................................48
Section 1100 Tables
Table 1. Buildup Test....................................................................................................................5
Table 2. DataFRAC Pressure Record ........................................................................................14
Table 3. Pressure Decline Analysis ............................................................................................18
Table 4. Input Data for FracNPV ................................................................................................25
Table 5. Output From the Inverse Module..................................................................................28
Table 6. Output of the PLACEMENT Simulator..........................................................................33
Table 7. Input Data for MLPP Production-Control Input .............................................................36
Table 8. Input Data for MLPP Production-Control Input .............................................................36

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 29 of 33

Table 9. Input Data for MLPP Production-Control Input............................................................. 37


Table 10. MLPP Output ............................................................................................................. 38
Table 11. Fracture Geometry..................................................................................................... 44
Table 12. Build-up Pressure Response ..................................................................................... 45
Section 1200 Appendix B - Fracturing Fluids
1 Introductory Summary................................................................................................................ 3
2 Water-Base Fluids ..................................................................................................................... 4
2.1 Polymers .............................................................................................................................. 4
2.1.1 Guar Gum................................................................................................................... 4
2.1.2 Hydroxypropylguar ..................................................................................................... 5
2.1.3 Hydroxyethylcellulose ................................................................................................. 6
2.1.4 Xanthan ...................................................................................................................... 7
2.1.5 Carboxymethylhydroxypropylguar .............................................................................. 8
2.2 Crosslinkers ......................................................................................................................... 8
2.2.1 Borate Crosslinker ...................................................................................................... 8
2.2.2 Organometallic Crosslinkers....................................................................................... 9
2.2.3 Crosslink Rate ............................................................................................................ 9
2.2.3.1 YF100 and YF200 Fluids ............................................................................... 10
2.2.3.2 YF300 and YF400 Fluids ............................................................................... 10
2.2.3.3 YF500 and YF600 Fluids ............................................................................... 10
3 Crosslinked Oil-Base Fluids ..................................................................................................... 11
3.1 YF"GO"III and YF"GO"IV Fluids......................................................................................... 12
4 Multiphase Fluids ..................................................................................................................... 12
4.1 Foams ................................................................................................................................ 12
4.2 Energized Fluids ................................................................................................................ 12
4.2.1 The Gas Phase......................................................................................................... 13
4.3 Emulsions .......................................................................................................................... 14
5 Acidic Fluids ............................................................................................................................. 15
6 Fracturing Fluid Characterization............................................................................................. 16
6.1 Rheology............................................................................................................................ 16
6.1.1 Shear and Temperature ........................................................................................... 16
6.1.2 Shear Rate ............................................................................................................... 16
6.1.3 Shear Stress ............................................................................................................. 16
6.1.4 Apparent Viscosity .................................................................................................... 16
6.1.5 Newtonian Fluids ...................................................................................................... 17
6.1.6 Non-Newtonian Fluids .............................................................................................. 17
6.2 Slurry Rheology ................................................................................................................. 19
6.3 Proppant Transport ............................................................................................................ 21
6.4 Fluid-Loss .......................................................................................................................... 22
6.5 Conductivity Damage from Fracturing Fluids ..................................................................... 23
6.5.1 The Effect of Water-Base Fracturing Fluids on Retained Permeability .................... 24
7 Fluid Selection ......................................................................................................................... 26

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 30 of 33

Section 1200 Figures


Fig. 1. The structure of guar. ........................................................................................................5
Fig. 2. The structure of HPG.........................................................................................................6
Fig. 3. The structure of HEC.........................................................................................................7
Fig. 4. The structure of Xanthan. ..................................................................................................7
Fig. 5. Borate crosslinking mechanism.........................................................................................9
Fig. 6. The structure of aluminum phosphate ester chain. .........................................................11
Fig. 7. Power-law exponent of a 40 lbm/1000 gal crosslinked water-base fluid. ........................17
Fig. 8. Consistency coefficient of a 40 lbm/1000 gal crosslinked water-base fluid. ....................18
Fig. 9. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid at 40 sec-1..........18
Fig. 10. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid at 170 sec-1.....19
Fig. 11. The effects of proppant on slurry viscosity of a Newtonian fluid....................................20
Fig. 12. The effects of proppant concentration on friction pressure of a water-base fluid. ........21
Fig. 13. Borehole fluid invasion zones........................................................................................23
Fig. 14. Effects of proppant concentration and porosity on postclosure polymer concentration.25
Fig. 15. Effect of polymer concentration on retained proppant-pack permeability......................26
Section 1200 Tables
Table 1. Comparison Of Nitrogen And Carbon Dioxide..............................................................13
Section 1300 Appendix C - Additives
1 Introductory Summary ................................................................................................................2
2 Fracturing Fluid Components .....................................................................................................2
2.1 Activators..............................................................................................................................3
2.2 Buffers ..................................................................................................................................3
2.3 Crosslinkers..........................................................................................................................3
2.4 Emulsifiers ............................................................................................................................3
2.5 Foaming Agents ...................................................................................................................4
2.6 Polymers...............................................................................................................................4
2.7 Potassium Chloride ..............................................................................................................4
3 Fracturing Fluid Additives ...........................................................................................................4
3.1 Bactericides ..........................................................................................................................4
3.2 Breakers ...............................................................................................................................5
3.2.1 Breakers for Water-Base Fluids ..................................................................................6
3.2.1.1 Enzyme Breakers .............................................................................................6
3.2.1.2 Oxidative Breakers ...........................................................................................6
3.3 Clay Stabilizers.....................................................................................................................9
3.3.1 Clay Types ..................................................................................................................9
3.3.2 Clay Control Methods................................................................................................12
3.3.2.1 Ionic Neutralization.........................................................................................12
3.3.2.2 Organic Barrier ...............................................................................................13
3.3.2.3 Particle Fusion................................................................................................13
3.4 Fluid-Loss Additives ...........................................................................................................13
3.5 Friction Reducers ...............................................................................................................14
3.6 Temperature Stabilizers .....................................................................................................15
3.7 Surfactants .........................................................................................................................17

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 31 of 33

3.7.1 Fluorocarbon Surfactants ......................................................................................... 22


3.7.2 Surfactant Selection.................................................................................................. 22
3.8 Nonemulsifying Agents ...................................................................................................... 23
3.8.1 Nonemulsifying Agent Selection ............................................................................... 23
4 Additive Selection..................................................................................................................... 25
Section 1300 Figures
Fig. 1. Apparent viscosity of 40 lbm/1000 gal crosslinked fluids with methanol and sodium
thiosulfate stabilizers....................................................................................................... 16
Fig. 2. Apparent viscosity of a 50 lbm/1000 gal crosslinked fluid containing sodium thiosulfate
and a 60 lbm/1000 gal crosslinked fluid containing methanol......................................... 16
Fig. 3. Surfactant orientation...................................................................................................... 18
Fig. 4. The wettability of oil/water/rock. ...................................................................................... 20
Section 1300 Tables
Table 1. Properties Of Common Dowell Surfactants ................................................................. 21
Table 2. The Effects of Wettability Change................................................................................ 21
Table 3. Summary of Surfactant Action on Mineral Surfaces .................................................... 22
Table 4. Properties Of Common Dowell Nonemulsifying Agents............................................... 24
Table 5. Additive Recommendation Guide................................................................................. 25
Table 6. Additive Selection Guide .............................................................................................. 26
Section 1400 Appendix D - Proppants
1 Introductory Summary................................................................................................................ 2
2 Physical Properties of Proppants ............................................................................................... 3
2.1 Proppant Strength................................................................................................................ 3
2.2 Grain Size and Grain-Size Distribution ................................................................................ 4
2.3 Quantities of Fines and Impurities ....................................................................................... 5
2.4 Roundness and Sphericity ................................................................................................... 6
2.5 Proppant Density ................................................................................................................. 6
3 Classes of Proppants ................................................................................................................. 6
3.1 Sand..................................................................................................................................... 6
3.2 Resin-Coated Proppants...................................................................................................... 6
3.2.1 Precured Resin-Coated Proppants............................................................................. 7
3.2.2 Curable Resin-Coated Proppants............................................................................... 7
3.2.3 Limitations Associated With Resin-Coated Proppants................................................ 7
3.2.3.1 Oxidative Breakers........................................................................................... 8
3.2.3.2 Borate-Crosslinked Fluids................................................................................ 9
3.2.3.3 Organometallic-Crosslinked Fluids .................................................................. 9
3.3 Intermediate-Strength Proppants......................................................................................... 9
3.4 High-Strength Proppants ..................................................................................................... 9
4 Conductivity ............................................................................................................................. 10
4.1 Closure Stress ................................................................................................................... 10
4.2 Embedment........................................................................................................................ 10
4.3 Fracture Width ................................................................................................................... 11
4.4 Proppant-Pack Porosity ..................................................................................................... 12

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Master Table of Contents Dowell
Page 32 of 33

4.5 Factors Operative in the Presence of Polymeric Fracturing Fluids ....................................13


5 Proppant Testing ......................................................................................................................13
6 Measurement of Proppant Addition to Fracturing Fluids ..........................................................14
7 Proppant Selection ...................................................................................................................16
8 Proppant Flowback...................................................................................................................16
Section 1400 Figures
Fig. 1. Strength comparisons of various types of proppants. .......................................................4
Fig. 2. The effect of feldspar contamination on conductivity.........................................................5
Fig. 3. The relationship of proppant concentration and fracture width (no embedment). ..........12
Section 1400 Tables
Table 1. Proppant-Pack Porosity Of Sand And Intermediate-Strength Proppant .......................12
Table 2. Density Table For Proppant Added To Fracturing Fluid ...............................................14
Section 1500 Appendix E - Fluid Loss
1 Introductory Summary ................................................................................................................2
2 Filtrate-Dependent Leakoff .........................................................................................................3
2.1 Wall-Building Coefficient.......................................................................................................5
2.2 Viscosity Control Coefficient.................................................................................................7
2.3 Compressibility Coefficient ...................................................................................................7
2.4 Fracturing Fluid Coefficient...................................................................................................8
2.5 Total Leakoff Volume............................................................................................................8
2.6 Fluid-Loss Mechanisms and Permeability ............................................................................8
3 Pressure-Dependent Leakoff......................................................................................................9
3.1 Geologic Discontinuities .......................................................................................................9
3.2 Opening Natural Fissures.....................................................................................................9
3.3 Fluid-Loss Control in Fissures ............................................................................................10
4 Types of Fluid-Loss Additives...................................................................................................11
4.1 Inert Particulates.................................................................................................................11
4.2 Soluble Particulates............................................................................................................11
4.3 Oil-in-Water Emulsions.......................................................................................................12
5 Formation Considerations ........................................................................................................12
5.1 Fluid Loss to the Rock Matrix .............................................................................................12
5.1.1 Pore-Size Determination ...........................................................................................13
5.1.1.1 Experimentally................................................................................................13
5.1.1.2 Calculation......................................................................................................13
5.2 Fluid Loss to Fissures.........................................................................................................13
6 In-Situ Measurement of Fluid-Loss Coefficients.......................................................................16
7 Guide to Dowell Fluid-Loss Additives .......................................................................................17
7.1 Fluid-Loss Additive Sizing ..................................................................................................18
Section 1500 Figures
Fig. 1. Zones of invasion. .............................................................................................................3
Fig. 2. The relationship of the fluid-loss coefficient to fluid volume and fracture length. .............4
Fig. 3. Typical fluid-loss data for a wall-building fluid. ..................................................................6
Fig. 4. Log-log plot of fracture pressure indicating possible fluid loss to fissures......................10

DOWELL CONFIDENTIAL
Section 000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Master Table of Contents
Page 33 of 33

Fig. 5. Opening of fissures. ........................................................................................................ 10


Fig. 6. Idealized G Plot (from the DataFRAC software). ............................................................ 16
Fig. 7. Bridging particle size versus pore-throat diameter.......................................................... 18
Fig. 8. Bridging particle size versus approximate permeability. ................................................. 19
Section 1500 Tables
Table 1. Influence Of Natural Fissures In Low-Permeability Rock............................................. 15
Table 2. Guide To Dowell Fluid-Loss Additives ......................................................................... 17
Section 1600 Appendix F - Equipment
1 Introductory Summary................................................................................................................ 2
2 Mixing and Blending Equipment................................................................................................. 4
2.1 PCM Precision Continuous Mixer ........................................................................................ 4
2.2 POD Blender ........................................................................................................................ 5
2.2.1 POD II Blender............................................................................................................ 7
3 Pumping Equipment................................................................................................................... 9
3.1 Pump Application Guidelines ............................................................................................... 9
3.1.1 Fracturing Fluid Viscosity............................................................................................ 9
3.1.2 Slurried Fluids Containing Proppant ........................................................................... 9
3.1.3 Proppant Concentration and Low Pump Speeds...................................................... 10
3.1.4 High Vapor Pressure Fracturing Fluids..................................................................... 10
3.1.5 Volumetric Efficiency ................................................................................................ 11
3.2 Nitrogen ............................................................................................................................. 11
3.3 Carbon Dioxide .................................................................................................................. 12
3.4 Pressure Multipliers ........................................................................................................... 12
4 Treating Equipment.................................................................................................................. 12
5 Sensors.................................................................................................................................... 13
6 Computing and Monitoring Equipment..................................................................................... 13
6.1 PACR Pumping, Acidizing, Cementing Recorder............................................................... 13
6.2 PPR Pumping Parameter Recorder ................................................................................... 13
6.3 FCS Computer System ...................................................................................................... 13
6.4 PAC Portable Acquisition Computer .................................................................................. 13
6.5 JMU Job Management Unit ............................................................................................... 14
7 Tools ........................................................................................................................................ 15
8 Support Equipment .................................................................................................................. 15
Section 1600 Figures
Fig. 1. Equipment positioning for a fracturing treatment (typical)................................................. 3
Fig. 2. The RampFRAC Service................................................................................................... 5
Fig. 3. The stairstep method of proppant addition. ...................................................................... 6

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 1 of 65

RESERVOIR EVALUATION

1 Introductory Summary............................................................................................................. 4

2 Well Performance..................................................................................................................... 5
2.1 Inflow Performance .............................................................................................................. 6
2.2 Tubing Intake ....................................................................................................................... 8

3 Well Test Interpretation ......................................................................................................... 10


3.1 Critical Variables ................................................................................................................ 11
3.2 Flow Regimes .................................................................................................................... 13
3.3 Boundary Effects................................................................................................................ 14
3.4 Diagnostic Plots ................................................................................................................. 15
3.5 Type Curves....................................................................................................................... 18
3.6 Computational System ....................................................................................................... 21
3.7 Steps for Analysis .............................................................................................................. 27
3.8 Example Analysis............................................................................................................... 30

4 Economic Analysis ................................................................................................................ 38


4.1 FracNPV Software ............................................................................................................. 40

5 Application ............................................................................................................................. 48

6 Eqquation Summary .............................................................................................................. 55


6.1 Oil IPR Equations............................................................................................................... 55
6.1.1 Darcy's Law ............................................................................................................... 55

6.1.2 Vogel Test Data ( Pr ≤ pb ) ......................................................................................... 55



6.1.3 Combination Vogel = Darcy Test Data ( Pr pb ) ....................................................... 55

6.1.4 Jones IPR .................................................................................................................. 57


6.2 Gas IPR Equations............................................................................................................. 58
6.2.1 Darcy's Law (Gas) ..................................................................................................... 58
6.2.2 Jones' Gas IPR (General Form) ................................................................................ 58

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 2 of 65

6.3 Backpressure Equation ......................................................................................................59


6.4 Transient Period Equations ................................................................................................60
6.4.1 Time to Pseudosteady State......................................................................................60
6.4.2 Oil IPR (Transient) .....................................................................................................60
6.4.3 Gas IPR (Transient) ...................................................................................................61
6.5 Completion Pressure Drop Equations ................................................................................61
6.5.1 Gravel-Packed Wells .................................................................................................61
6.5.2 Open Perforation Pressure Drop ...............................................................................63

FIGURES

Fig. 1. Pressure losses in complete systems (after Mach, Proano and Brown). ..........................5
Fig. 2. Location of various nodes (Mach et al., 1981)...................................................................6
Fig. 3. Typical IPR curve. .............................................................................................................7
Fig. 4. Vogel's composite IPR.......................................................................................................8
Fig. 5. Vertical multiphase flow: How to find the flowing bottomhole pressure. ...........................9
Fig. 6. Tubing intake curve. ..........................................................................................................9
Fig. 7. Deliverability of the producing system. ............................................................................10
Fig. 8. Well log............................................................................................................................12
Fig. 9. Well log............................................................................................................................12
Fig. 10. Radial flow. ....................................................................................................................13
Fig. 11. Linear flow in the formation. ..........................................................................................14
Fig. 12. Bilinear flow. ..................................................................................................................14
Fig. 13. Conditions associated with the boundries. ....................................................................15
Fig. 14. Pressure change and elapsed time to use in a drawdown. ...........................................16
Fig. 15. Log-Log plot. .................................................................................................................17
Fig. 16. Pressure change and elapsed time. ..............................................................................17
Fig. 17. Complete log-log behavior.............................................................................................17
Fig. 18. This is a reproduction from a type curve described in World Oil, (Oct. 1983). ..............19
Fig. 19. Series of pressure, pressure derivative, and specialized plots for common
reservoir features..........................................................................................................20
Fig. 20. Diagnostic log-log plot. ..................................................................................................21
Fig. 21. Two Horner plots. ..........................................................................................................22
Fig. 22. Model-Verified Interpretation. ........................................................................................23
Fig. 23. Conceptual model catalog. ............................................................................................24
Fig. 24. NODAL plot. ..................................................................................................................25
Fig. 25. Sequence simulation. ....................................................................................................25
Fig. 26. Simulated validation. .....................................................................................................26
Fig. 27. PVT plot.........................................................................................................................26
Fig. 28. Matching a diagnostic log-log plot to a type curve.........................................................29
Fig. 29. Matching a diagnostic log-log plot to a type curve.........................................................29
Fig. 30. Log-Log Plot ..................................................................................................................30
Fig. 31. Generated type curve with the log-log diagnostic match...............................................32

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 3 of 65

Fig. 32. Semilog presentation using a superposition type curve and the data points
from the buildup. .......................................................................................................... 33
Fig. 33. Cartesian plot of the simulated pressure and the actual measured data...................... 33
Fig. 34. Decline curve and sensitivity plot. ................................................................................. 34
Fig. 35. Decline curve and sensitivity plot. ................................................................................. 34
Fig. 36. Sensitivity plot. .............................................................................................................. 35
Fig. 37. Sensitivity plot. .............................................................................................................. 35
Fig. 38. Sensitivity plot. .............................................................................................................. 36
Fig. 39. Sensitivity plot. .............................................................................................................. 36
Fig. 40. Sensitivity plot. .............................................................................................................. 37
Fig. 41. Plot of the transient IPR curves with the tubing intake and wellhead
pressure of 875 psi....................................................................................................... 37
Fig. 42. Transient IPR plot for the same tubing, but using different wellhead pressures to
generate the plot shown in Fig. 32. .............................................................................. 38
Fig. 43. Conceptual NPV calculation. Case A: revenue is larger than the cost, resulting in a
positive NPV; Case B: revenue is less than the cost, resulting in a negative NPV. ..... 40
Fig. 44. Components of the NPV calculation. ............................................................................ 41
Fig. 45. Fracture in a bounded reservoir.................................................................................... 41
Fig. 46. Constant-rate type curve for finite-conductivity fracture  closed square system
(xe/ye = 1). ...................................................................................................................42
Fig. 47. Constant-rate type curve for finite-conductivity fracture  closed rectangular system
(xe/ye = 2). ...................................................................................................................42
Fig. 48. Constant-rate type curve for finite-conductivity fracture  closed rectangular system
(xe/ye = 4). ...................................................................................................................43
Fig. 49. Transient IPR for a fracture well in a closed square reservoir. ..................................... 44
Fig. 50. Tubing intake curve....................................................................................................... 45
Fig. 51. FracNPV analysis for one, two and three years versus fracture length. ....................... 46
Fig. 52. Cumulative production versus fracture length............................................................... 46
Fig. 53. Production decline versus time. .................................................................................... 47
Fig. 54. Cumulative production versus time. .............................................................................. 47
Fig. 55. Well performance tracking form. ................................................................................... 51
Fig. 56. Well performance tracking form. ................................................................................... 53

TABLES

Table 1. Results of Buildup Test ................................................................................................ 31


Table 2. Maximum Fracture Lengths for Drainage Shapes ....................................................... 43

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 4 of 65

1 Introductory Summary
Evaluation is a critical part of any process where understanding and improvement
are goals. Understanding and improving the fracturing treatments demand that the
best information be available for design, and that the results obtained from the
execution of the design be analyzed.
The appropriate fracturing treatment for a given well has been hard to design
because of the numerous variables involved. The evaluation of fracturing treatments
has been difficult because many of these same design variables are also used for
production simulation. The use of inaccurate reservoir variables to design
treatments therefore leads to poor production estimates. Many times a production
estimate is not even made, and evaluation then ceases to be a part of the
technology required to supply a value to the client.
A successful fracturing treatment has too long been defined as “one that was
pumped without problems.” A successful treatment needs to be defined as “one that
provides the production predicted by the design process.” A fracture stimulation
treatment should optimize production. Most of the time the treatment does not
optimize production because realistic values for the critical variables used in the
design process have not been properly identified. Critical variables are those
variables that have the greatest impact on the production obtained from a fracturing
treatment. Permeability, for example, is a critical variable. An accurate value for
these variables (rather than general estimates) is very important for a realistic design
and for accuracy in predicting the production response. Optimization is possible with
proper identification and use of the accurate critical variables.
The critical variables used to optimize stimulation treatments can be placed into
three categories. Each category contains variables needed to determine specific
design criteria. These three categories are (1) reservoir and producing system
variables for determining the production response from the well, (2) stimulation
design variables that determine achievable fracture geometry, and (3) economic
variables that determine the optimum treatment. Reservoir evaluation deals
primarily with the reservoir and producing system variables and with the economic
variables. Design variables are covered in more detail in other sections of this
manual.
Reservoir and producing system variables are those variables obtained from well
performance testing/analysis and from analyzing surface and wellbore plumbing
configurations. Economics optimization deals with the production, net discounted
production revenue and cost of the fracturing treatment.
The teaching of the total technology involved in pressure transient and well
performance analysis is beyond the scope of this section. The user should study the
applicable references and attend training sessions on well performance testing
whenever possible. Therefore, certain assumptions must be made and the section
will deal with how to apply well performance data to fracture design and evaluation.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 5 of 65

2 Well Performance
Well performance can be defined simply as the ability of a well to produce reservoir
fluids to the surface by either natural flow or by artificial lift. The physical description
of a typical well is shown in Fig. 1. The figure also illustrates the pressure losses
that can occur from the reservoir to the separator. A node is any point in the
production system between the drainage boundary and the separator where the
pressure can be calculated as a function of the flow rates. Fig. 2 shows the location
of various nodes in a producing system. It is important to understand the
components of the production system, because different pressure-loss relationships
are used in an analysis method for designing and optimizing the total system
(NODAL* analysis).

Fig. 1. Pressure losses in complete systems (after Mach, Proano and Brown).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 6 of 65

Fig. 2. Location of various nodes (Mach et al., 1981).

2.1 Inflow Performance


The Inflow Performance Relationship (IPR) is defined as the functional relationship
between the production rate and the bottomhole flowing pressure. The IPR is
defined in the pressure range between the average reservoir pressure and
atmospheric pressure. The flow rate corresponding to the atmospheric pressure at
bottomhole is defined as the absolute open flow potential (AOFP). A typical IPR for
a single-phase liquid is shown in Fig. 3. Darcy's law for radial flow is used to obtain
the flow rates necessary to construct this IPR. Using re = 1466 ft, rw = 0.583 ft,
st = 0 and no turbulence, Darcy's law simplifies to

qo =
kh
µ o Bo ( pr − pwf ) (k in Darcy) (1)

This simple equation is often used for the estimation of flow rates from oil wells.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 7 of 65

Fig. 3. Typical IPR curve.


It was stated that the IPR shown in Fig. 3 is a straight-line relationship based on
Darcy's law, and the AOFP is the maximum flow rate with the atmospheric pressure
at the bottomhole. The Productivity Index (PI) is the absolute value of the slope of
the IPR straight line. Therefore, a simple equation for PI can be written as follows.
q
PI = (2)
( pr − pwf )
The two equations are shown to illustrate the basic simplicity in constructing and
using the IPR, as well as to point out a reminder or caution when dealing with IPR
curves. The PI concept is not used for gas wells because the rate in Darcy's
equation for gas is a function of pressure squared. Consequently, the IPR becomes
a curve rather than a straight line, and the slope therefore changes with the rate in
this case. The PI concept cannot be used in two-phase systems (gas/liquid),
because Darcy's equation can only be used to construct the portion of the line where
the pressure is above the bubblepoint (pb) for a single-phase liquid system. The
portion of the curve below the bubblepoint pressure must be corrected (Vogel's
equation) for gas. Fig. 4 shows an example of a composite IPR and how the
maximum production would be overestimated if Vogel's IPR was not applied.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 8 of 65

Fig. 4. Vogel's composite IPR.


Vogel's IPR does not consider a skin factor, and therefore is only applicable for
undamaged wells. Standing's extension of Vogel's IPR uses the concept of a flow
efficiency factor (FE) to extend the effect of skin on Vogel's equation. The reader
should examine the method Standing used, and also understand the concept of skin
factors and the impact skin has on the producing system.

2.2 Tubing Intake


The previous topic of inflow performance developed the concept of using Darcy's
equation, and modifications, to construct IPR plots. Obviously, the characteristics of
an IPR curve are sensitive to the variables used in the equations, and each change
in any variable will produce a new IPR curve. The nodes examined so far deal with
the actual reservoir variables. This new topic deals with the effect that the tubular
configuration has on the production.
Gradient curves are used to generate a plot of rate versus bottomhole flowing
pressure. Gradient curves are available for a wide range of flowline or tubing size
and a set of fixed flow and fluid parameters. Fig. 5 is an example gradient curve for
2-7/8-in. tubing and 1000 BPD of liquid production at 50% oil. A particular set of
conditions is chosen (wellhead pressure, tubing diameters, and fluid type). The
gradient curve is then used to obtain different rates versus bottomhole flowing
pressure, and the tubing intake curve is constructed as shown in Fig. 6.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 9 of 65

Fig. 5. Vertical multiphase flow: How to find the flowing bottomhole pressure.

Fig. 6. Tubing intake curve.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 10 of 65

Additional comments concerning Fig. 6 are:


• Negative slope at low rate is indicative of unstable flow in the pipe.
• Inflection point in the curve is the critical rate below which the gas will slip by the
liquid and the well will load up. Any rate below the critical rate will kill the well.
• Positive-slope portion of the curve shows the conditions where the velocity is high
enough to move the fluids to the surface.
The tubing intake curve plotted with the IPR curve is shown in Fig. 7. The
intersection of the two curves determines the deliverability of the producing system.

Fig. 7. Deliverability of the producing system.


Additional information is available in the references, which cover in more detail the
nodes that can be analyzed similar to the intake curve. Perforation size and density,
for example, can have a major impact on the production.

3 Well Test Interpretation


Reservoirs not only differ in lithology but also in behavior. Testing in many different
reservoirs has shown that the possible number of behaviors from a well test is limited
to three; therefore, only a few interpretation models are required for analysis. The
three main types of behavior are (1) homogeneous, (2) dual porosity, and (3) dual
permeability.
1. Homogeneous Behavior. There is only one porous medium producing into the
well. This is an overall definition (mathematical) and does not actually mean that
the reservoir is homogeneous, but that there is a characteristic shape of the
reservoir pressure response to a change in rate.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 11 of 65

2. Dual-Porosity Behavior. Two homogeneous porous media are interacting. Each


medium is of distinct porosity and permeability and can be uniformly distributed,
or they can be segregated. However, only one medium produces fluid into the
well and the other medium acts as a source. Dual-porosity behavior may be
used to explain naturally fractured or fissured reservoirs
• single-layered reservoirs, but with high-permeability variation through the
reservoir thickness
• reservoirs that are multilayered with a high-permeability contrast between the
layers.
3. Dual-Permeability Behavior. Two distinct porous media, each producing into the
well. This behavior may be described as multilayered reservoir with low-
permeability contrast
• multiple zones separated by impermeable layers
• partial completion.
These three types of behavior have certain characteristics that are recognizable on
various diagnostic plots that will be discussed later.

3.1 Critical Variables


There are three critical variables obtained from well test interpretation, (1) reservoir
pressure, (2) skin, and (3) permeability. Permeability can be obtained from core
analysis, but unless the analysis is performed under in-situ conditions the
permeability could be off several orders of magnitude too high. Permeabilities
obtained from pressure tests are considered the most reliable for true reservoir
permeability because these tests are a direct measure of the in-situ reservoir
response. Because the correct interpretation of the pressure test data is critical for
accurate permeability and flow-rate predictions, it is important to note and keep in
mind the following.
• All well-test-interpretation equations are solved in terms of kh/(µβ)
(transmissibility). Therefore, an accurate calculation of k depends on accurate
values for reservoir thickness, fluid viscosity, and formation volume factor.
• All well test interpretations make the assumption that there is a major producing
phase, either oil, oil and water (total liquid) or gas. Values for the viscosity and
formation volume factor are based on the major phase of production. For a well
producing oil and water, a weighted average viscosity and weighted average
volume factor should be used to calculate k.
Remember that h is the net productive interval and not just the perforated height (hp).
The net interval thickness is also different from the gross thickness (hg) normally
used as height growth when designing or evaluating hydraulic fracturing. Fig. 8 and
Fig. 9 are examples of well logs showing the possible combinations of net,
perforated and gross height.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 12 of 65

Fig. 8. Well log.

Fig. 9. Well log.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 13 of 65

3.2 Flow Regimes


Pressure transient analysis is used to analyze the pressure change with time under
the constant producing rate or constant bottomhole pressure. Wells can be
unstimulated, acidized or stimulated with hydraulic fractures that are either etched or
propped. Stimulation alters part of the producing system and can dramatically
increase the production. These different well conditions have different flow regimes
that affect the pressure behavior at the well. The following flow regimes are
important for the purpose of analysis.
• Radial Flow. Flowlines at all elevations in the porous zone around the wellbore
are radial. Radial flow converges in producing wells and diverges in the case of
injection wells. Fig. 10 illustrates this regime.
• Linear Flow. Wells that have been hydraulically fractured can show both linear
flow from the matrix to the fracture and linear flow along the fracture to the
wellbore. The dominant flow for a well with a very high-conductivity fracture will
be from the matrix to the fracture. Fig. 11 illustrates linear flow.
• Bilinear Flow. The flow of fluid in a well with a finite-conductivity fracture depends
not only on the linear flow from the matrix but also the linear flow along the
fracture. The bilinear flow regime is shown in Fig. 12.

Fig. 10. Radial flow.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 14 of 65

Fig. 11. Linear flow in the formation.

Fig. 12. Bilinear flow.


Fluid flow in porous media is expressed mathematically by the diffusivity equation.
This equation is a partial differential equation and is solved with specific inner
(wellbore) and outer boundary conditions.

3.3 Boundary Effects


Transient tests in wells show that the pressure response is affected by boundary
conditions as well as by the basic reservoir behavior such as homogeneous, dual
porosity and dual permeability. Fig. 13 shows the conditions associated with the
boundaries. Some additional comments are made to clarify outer boundaries.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 15 of 65

Fig. 13. Conditions associated with the boundaries.


• Constant pressure outer boundary assumes the reservoir has an active water
drive or pressure maintenance program (fluid injection) that maintains constant
reservoir pressure (steady state). This also means that fluid produced from the
well will be replaced at the same rate by fluid flowing across the boundary.
• A closed outer-boundary reservoir can never reach steady state because fluid is
not replaced. This no-flow outer boundary can be set by impermeable barriers,
faults, pinchout, or other producing wells across the boundary. The transient
pressure response does reach the outer boundary at late times when the rate of
pressure change with time is constant. This is called the pseudosteady state.
The time needed to reach the pseudosteady state is dependent on the geometry
(shape factor) of the reservoir.

3.4 Diagnostic Plots


Pressure transient analysis is basically a pattern-recognition process and involves
plotting the pressure change versus time. The diagnostic plot is the first plot which
must be made in any analysis. This is a log-log plot (using all of the available data)
of pressure function change versus time during the period to be analyzed. The log-
log scale is used because it allows the characteristic shape of a response to be
observed without distortion due to the range of data. The diagnostic log-log plot is
used to
• select the basic interpretation model
• identify different flow regimes, such as the basic reservoir, inner boundary and
outer boundary

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 16 of 65

• calculate the parameters from the comparison of the log-log curve and type
curves.
A pressure function must be selected when generating a log-log plot. Use the direct
pressure change for oil wells, and pressure squared for gas wells. Real gas
pseudopressure, M(p), is preferred for gas wells when computer software is used.
See the references for a more detailed discussion on pressure functions.
Analysis can be made using drawdown data as well as buildup data. Fig. 14 shows
the pressure change and elapsed time to use in a drawdown. The log-log plot will
appear as shown in Fig 15. Fig. 16 shows the pressure change and elapsed time to
use when a log-log plot is made to analyze buildup data. The pressure response of
any period is affected by the prior periods, and therefore pressure transient analysis
of a buildup must take into account the effects of the previous flow period. The
transient effects of the flow period continue to affect the pressure in the reservoir
after the well is shut in. This effect is not indicated on the pressure recorder. The
complete log-log behavior is illustrated in Fig. 17, which shows a well with wellbore
storage in a closed homogeneous reservoir. Specialized plots can be used to
confirm the flow regime identified on the diagnostic log-log plot. Details of the
specialized plots will not be covered, but examples will be shown later.

Where: pwf (∆t) = flowing pressure at ∆t,


pi = initial pressure,
to = time when the well was opened, and
∆p(∆t) = pi - pwf (∆t)
Fig. 14. Pressure change and elapsed time to use in a drawdown.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 17 of 65

Fig. 15. Log-log plot.

Fig. 16. Pressure change and elapsed time.


Note: This is the pressure change and elapsed time to use when a log-log plot is
made to analyze buildup data.

Fig. 17. Complete log-log behavior.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 18 of 65

3.5 Type Curves


Type curves are graphical representations of the solution of the diffusivity equation
for the constant rate drawdown under different boundary conditions. Type curves
are available for many reservoir and well types. Some of these type-curve
combinations are
• homogeneous reservoir with or without wellbore storage and skin
• homogeneous reservoir with or without induced fractures in the wellbore
• dual porosity or naturally fractured reservoirs
• layered reservoir.
The three variables in the x, y, and z dimensions in the type curve are dimensionless
pressure, dimensionless time and a variable representing either the near wellbore
condition or the boundary shape. Depending on the wellbore condition, the z
variable may be (1) wellbore storage, c, and skin, s, in the case of homogeneous
reservoirs or (2) fracture conductivity, CfD, in the case of wells with an induced
fracture. Fig. 18 is a type curve for a homogeneous reservoir with wellbore storage
and skin. The advantage of this type curve is the ease of matching and clear
definition of the flow regimes. The dimensionless variables for this type curve are
shown on the upper portion of the curve. The presentation of these dimensionless
variables in log-log coordinates makes it possible to match the pressure versus time
data obtained from a well test.
The use of a pressure derivative represents a major advancement in pressure
transient analysis. The derivative is useful because not only the pressure curve but
also the pressure derivative curve must match the analytical solution. Pressure
trends can be confusing at middle and late times and subject to many interpretations,
but the derivative values are much more definitive. Fig. 19 is a series of pressure,
pressure derivative and specialized plots for common reservoir features.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 19 of 65

Fig. 18. This is a reproduction from a type curve described in World Oil, (Oct. 1983). The
original graph is in three colors.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 20 of 65

Fig. 19. Series of pressure, pressure derivative, and specialized plots for common
reservoir features.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 21 of 65

3.6 Computational System


Mistakes based on incorrect well test analysis can be embarrassing and costly. It
has become too time consuming to process well test data by hand, and the margin of
error is too high. Therefore, a more economical and practical analysis system is
necessary.
The STAR* (Schlumberger Transient Analysis and Report) software incorporates the
essentials of modern reservoir engineering. This comprehensive computational
system is easily capable of handling the most complex transient analysis. The
software is menu driven, and correct results can be obtained with a minimum of
training. The system is capable of performing sensitivity analysis (NODAL analysis),
validation and interpretation. The following is an example of the difficulty in making
an interpretation, and how the derivative and a model to verify the interpretation are
tremendous advantages for well test interpretation.
A 30-hr buildup test was conducted on a Mid-Continent oil well. Fig. 20 is the
diagnostic log-log plot indicating that the Horner analysis straight line might be
difficult because of the changes indicated at Point A and Point B.

Fig. 20. Diagnostic log-log plot.


Fig. 21 shows two Horner plots, one with the straight line through the points at A and
the other through the points at B. The permeability and skin for each interpretation
are shown. The question now becomes “which is the correct Horner straight line?”
The answer is neither.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 22 of 65

Fig. 21. Two Horner plots.


Fig. 22 is the Model-Verified Interpretation of the test data, and indicates the
presence of a heterogeneous system. A two-porosity model with decreasing
wellbore storage matches the entire buildup, and note the difference in
permeabilities and skin. Anyone can draw a Horner straight line, but it must be the
correct straight line. Fig. 19 is by no means all encompassing, but is there a similar
type pattern to that shown in Fig. 22.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 23 of 65

Fig. 22. Model-verified interpretation.


The STAR software can be used to actually suggest the proper type of test, and
select the test objectives that are directly related to well productivity. The first step is
to select a conceptual model most closely resembling the well, and match it with the
petrophysical parameters of each productive zone derived from the log and
geological data. Fig. 23 shows the conceptual model catalog of the program.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 24 of 65

Fig. 23. Conceptual model catalog.


The NODAL analysis part of the software can be used to include a comparison of the
completion variables and, at the same time, study the economics of the completion
options. Fig. 24 shows an example NODAL plot of some selected completion
options. The actual test sequence can now be simulated. This simulation is a two-
part process best described as a series of flow-rate variations and flow periods
necessary to achieve the desired results. The test sequence can be simulated within
the software to allow the estimation of the ranges of values that will be encountered
in the test. Validation of the design is accomplished by simulating the interpretation
to verify that the test will determine the targeted objectives. Fig. 25 shows an
example simulation of a test sequence and expected response. Fig. 26 is the
second part, and is an example of the simulated validation showing the estimation of
the times for essential test features such as the end of the wellbore storage, onset of

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 25 of 65

boundary conditions, etc. Fluid-phase behavior can now be considered at this time
to ensure that the tests are not unnecessarily complicated by phase changes.
Fig. 27 is an example of a simulated PVT plot. The software calculates all fluid
properties, and the option is available to input actual PVT data. In many cases, the
calculated values compare very close to the values from real PVT data on given well.

Fig. 24. NODAL plot.

Fig. 25. Sequence simulation.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 26 of 65

Fig. 26. Simulated validation.

Fig. 27. PVT plot.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 27 of 65

3.7 Steps for Analysis


Analysis requires the well test data, well and formation data, and fluid data. Well test
data required are
• Pressure (p)
• Rates (q)
• Time (t).
Well and formation data required are
• Depth (at midpoint of perforations)
• Tubing Size
• Wellbore Radius (bit size) (rw)
• Formation Thickness (net producing interval)
• Perforated Thickness
• Perforation Size and Density
• Formation Temperature (T)
• Porosity (fraction).
Fluid data required are
• Liquid (oil)gravity (either specific or API)
− gas/oil ratio (GOR)
− bubblepoint pressure (pb)
− formation/volume factor (βo)
− viscosity (µ)
− compressibility (co)
− solution gas/oil ratio (Rs).
• Gas
− specific gravity
− formation/volume factor (βg)
− critical pressure (pc)
− critical temperature (Tc)
− gas deviation factor (Z)
− viscosity (µ)
− compressibility (cg).

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 28 of 65

Assuming all of the data are available, six steps should be completed for a
successful analysis. The STAR software performs these steps very quickly, and may
or may not follow the exact sequence.
1. Draw the log-log diagnostic plot. Use the type curve that will be used and lay
tracing paper over the type curve. Trace the main grids and mark the scales.
Plot the pressure and time data on the tracing paper but do not draw a line
through the points.
2. Identify the flow regimes. For example, the early time will exhibit the
characteristic shapes (usually one dominates) such as wellbore storage (unit
slope), finite-conductivity fractures (quarter slope) or high-conductivity fractures
(half slope). Refer again to Fig. 13 to see the characteristics that can be
associated with the middle and late time.
3. Specialized plots (Cartesian or semilog) may now be drawn using only those
individual data points that indicated the special shape on the diagnostic log-log
plot. Wellbore storage, for example, is a unit slope and should therefore be a
straight line through the origin on a Cartesian plot. Finite-conductivity fractures
are quarter slope and a Cartesian plot of the pressure versus the fourth root of
the time will also be a straight line passing through the origin. The references
show many examples of these plots.
4. Calculate the parameters from the specialized plots. Caution is recommended
and a reminder that accurate parameters must be obtained for the analysis to be
beneficial. It is again recommended that the beginner study the references and
have access to the STAR software as well as someone that can explain the
software.
5. Select the most appropriate type curve based on the previous analysis showing
the most likely reservoir behavior and well type. Place the diagnostic log-log plot
over the type curve and shift it to a position which gives a best fit. The axis of the
two graphs must be parallel at all times. Select a match point. This match point
can be anywhere, but choose a point that is easy to identify such as the intersect
of the major axis. Record the match point coordinates of the diagnostic plot
(pressure and time) and the dimensionless values of the underlying type-curve
grid (tD and pD). These values can now be used to calculate the desired
parameters using the equations given in Fig. 18. An example of how to match a
diagnostic log-log plot to a type curve is shown using Fig. 28 and Fig. 29. Make a
copy of each of these and then use a light table or window to practice moving the
plot to get the match shown, while keeping the axis parallel.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 29 of 65

Fig. 28. Matching a diagnostic log-log plot to a type curve.

Fig. 29. Matching a diagnostic log-log plot to a type curve.


6. Check for the consistency of the results. Numerical results from specialized
analysis and from the log-log type-curve analysis should agree with less than
10% error. Be sure that the data points used to identify the flow regime on the
specialized plots also match the same flow regime on the type curve. Verify that
the straight line has been drawn through the correct points. Verify that the right
match point was selected. If a consistent result cannot be obtained, then it is
possible that the wrong model was selected.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 30 of 65

3.8 Example Analysis


This example shows the steps as well as an example of an analysis made using the
STAR software. This is a real well test from a gas well that had been fracture
stimulated (with proppant) several years prior to the buildup. Real PVT data were
not available so the calculated data were used. Additional well parameter data are
• total compressibility (1/psi) = 2.098E-04
• gas gravity = 0.65
• liquid/Gas Ratio (STB/MMscf) = 1.90
• viscosity (cp) = 0.01946
• porosity (%) = 8.6
• reservoir temperature (°F) = 275
• hole size (in.) = 7.875
• net pay (ft) = 31
The buildup test was successful (see data in Table 1), and the data were first
analyzed by a log-log diagnostic shown in Fig. 30. The log-log plot is laid over the
type curve for a preliminary match that just serves as the starting point for the
analytical solver in the software. The solver is a numerical simulator and obtains a
unique match of the well test data.

Fig. 30. Log-log plot.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 31 of 65

Table 1. Results of Buildup Test


TEST PHASE: SHUTIN PERIOD # 1 FINAL FLOW PRESSURE - 1461 PSIA
PRODUCING TIME - 48930 HR

TIME DATE ELAPSED DELTA BOT HOLE DELTA P LOG


OF DAY PRESSURE HORNER
HH:MM;SS DD-MMM TIME, HR TIME, HR PSIA PSI TIME
12:00:01 9 - MAR 0.000 0.000 1461 0
12:04:48 9 - MAR 0.080 0.080 1729 268 5.7874
12:30:00 9 - MAR 0.500 0.500 1752 291 4.9908
12:49:48 9 - MAR 0.830 0.830 1801 341 4.7706
13:09:36 9 - MAR 1.160 1.160 1823 362 4.6252
13:30:00 9 - MAR 1.500 1.500 1837 376 4.5135
13:49:48 9 - MAR 1.830 1.830 1848 387 4.4272
14:00:00 9 - MAR 2.000 2.000 1854 393 4.3886
14:15:00 9 - MAR 2.250 2.250 1860 399 4.3374
14:30:00 9 - MAR 2.500 2.500 1867 406 4.2917
14:45:00 9 - MAR 2.750 2.750 1872 411 4.2503
15:00:00 9 - MAR 3.000 3.000 1878 417 4.2125
15:15:00 9 - MAR 3.250 3.250 1880 420 4.1777
16:30:00 9 - MAR 4.500 4.500 1903 442 4.0364
17:30:00 9 - MAR 5.500 5.500 1912 451 3.9493
18:30:00 9 - MAR 6.500 6.500 1923 462 3.8767
20:00:00 9 - MAR 8.000 8.000 1936 476 3.7866
21:00:00 9 - MAR 9.000 9.000 1946 486 3.7354
22:00:00 9 - MAR 10.000 10.000 1952 491 3.6897
23:00:00 9 - MAR 11.000 11.000 1960 499 3.6483
0:00:00 10 - MAR 12.000 12.000 1968 507 3.6105
1:00:00 10 - MAR 13.000 13.000 1972 512 3.5758
2:00:00 10 - MAR 14.000 14.000 1979 519 3.5436
4:00:00 10 - MAR 16.000 16.000 1988 527 3.4856
6:00:00 10 - MAR 18.000 18.000 2000 540 3.4345
8:00:00 10 - MAR 20.000 20.000 2009 549 3.3887
10:00:00 10 - MAR 22.000 22.000 2018 558 3.3474
12:00:00 10 - MAR 24.000 24.000 2027 566 3.3096
16:00:00 10 - MAR 28.000 28.000 2041 580 3.2427
20:00:00 10 - MAR 32.000 32.000 2053 592 3.1847
0:00:00 11 - MAR 36.000 36.000 2068 607 3.1336
4:00:00 11 - MAR 40.000 40.000 2078 618 3.0879
8:00:00 11 - MAR 44.000 44.000 2091 630 3.0465
12:00:00 11 - MAR 48.000 48.000 2104 644 3.0088
20:00:00 11 - MAR 56.000 56.000 2123 662 2.9419
4:00:00 12 - MAR 64.000 64.000 2144 683 2.8840
12:00:00 12 - MAR 72.000 72.000 2159 698 2.8329

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 32 of 65

Fig. 31 is the generated type curve with the log-log diagnostic match. The best-fit
type curve was a homogeneous, finite-conductivity vertical fracture with variable
wellbore storage. Note that all of the critical parameters have been calculated at this
point in the process.
Two model-verification plots were then produced. Fig. 32 is a semilog presentation
using a superposition type curve and the data points from the buildup. Fig. 33 shows
the other model verification using a Cartesian plot of the simulated pressure and the
actual measured data.
Production decline curves were then generated to predict the future production of the
zone (1) under current conditions, and (2) if the liquid/gas ratio were changed.
NODAL analysis was used to generate these two decline curves (Fig. 34 and
Fig. 35) as well as the sensitivity plots that follow.
Fig. 36 is a sensitivity plot showing the change in the production with the change in
the wellhead pressure.
The analysis then evaluated what the new production would be if the well were
refractured. A fracture conductivity of 1400 md-ft was input to generate the
sensitivity analysis of production, time and fracture length. Fig. 37 through Fig. 40
show these sensitivity plots.

Fig. 31. Generated type curve with the log-log diagnostic match.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 33 of 65

Fig. 32. Semilog presentation using a superposition type curve


and the data points from the buildup.

Fig. 33. Cartesian plot of the simulated pressure and the actual measured data.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 34 of 65

Fig. 34. Decline curve and sensitivity plot.

Fig. 35. Decline curve and sensitivity plot.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 35 of 65

Fig. 36. Sensitivity plot.

Fig. 37. Sensitivity plot.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 36 of 65

Fig. 38. Sensitivity plot.

Fig. 39. Sensitivity plot.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 37 of 65

Fig. 40. Sensitivity plot.


Fig. 41 is the plot of the transient IPR curves with the tubing intake and wellhead
pressure of 875 psi. Fig. 42 is the transient IPR plot for the same tubing, but using
different wellhead pressures to generate the plot shown in Fig. 32.

Fig. 41. Plot of the transient IPR curves with the tubing intake and
wellhead pressure of 875 psi.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 38 of 65

Fig. 42. Transient IPR plot for the same tubing, but using different wellhead
pressures to generate the plot shown in Fig. 32.

4 Economic Analysis
Up to now this study has dealt primarily with evaluating the reservoir and plumbing
system to obtain values for the critical variables, and to allow an accurate estimation
of production at the surface where the production is actually measured and sold.
The example analysis in the STAR software showed that hydraulic fracturing should
also be evaluated because of the impact fracturing has on the reservoir. Not only
does the reservoir deliverability and producing system need consideration, but also
the fracture mechanics, fracturing fluid characteristics, proppant and transport of the
proppant, operational constraints and economics. All of these considerations must
be integrated to produce the most cost-effective design, and to maximize the
benefits of a fracturing treatment. Maximizing these benefits requires a balance
between the fracture characteristics and reservoir properties to optimize the
reservoir deliverability. The final fracture design must be achievable from the
execution approach, and the cost of the treatment versus the value obtained must be
considered for true optimization.
Value is critical because money not only has a current purchasing value, but it also
has a time value. Even if risk and inflation are not considered, a dollar today is worth
more than the same dollar a year from now because the dollar can be invested
during the year. When the dollar is invested, it earns interest. Expressing the
concept very simply, the present value (PV) of an amount is the time zero value of a
future cash flow (FV) discounted N years at a % interest rate (i) per period. A simple
equation expressing this and the Net Present Value (NPV) is as follows.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 39 of 65

FV
PV = (3)
(1 + i ) n

N
Discounted Well Revenue = Σ
n= 1
(4)
Total Annual Net Revenue During Yr n
(1 + i )n
Total Net Revenue (TNR) $/BBL or MCF X (frac - nonfrac) production. The fracture
NPV is then expressed as
Fracture NPV = Discounted Well Revenue - Treatment Cost.
Many companies have software to evaluate economics through the NPV as it relates
to fracturing. However, it is doubtful if very many have the entire analysis package.
FracNPV analysis uses the fracture NPV to select the optimum propped fracture
treatment for a specific well and reservoir conditions. The software couples the
reservoir response, well hydraulics, fluid rheology, fluid volume and proppant
concentration, pumping parameters, rock properties, and closure stress value. A
range of fracture lengths is used and the economics are applied using the NPV
concept. The rate of return can be misleading when dealing with incremental
revenues; therefore, the optimum fracture size is defined as the one that provides
the maximum NPV. Fig. 43 illustrates the NPV concept with two cases. Case A
shows that the incremental revenue far exceeds the cost over the fracture lengths
shown. Subtracting the cost from the revenue results in the characteristic bell-
shaped curve of the NPV. Some wells may not be good candidates for fracturing,
and Case B shows an example where the incremental revenue does not justify the
cost because the NPV curve is negative.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 40 of 65

Fig. 43. Conceptual NPV calculation. Case A: revenue is larger than the cost, resulting
in a positive NPV; Case B: revenue is less than the cost, resulting in a negative NPV.

4.1 FracNPV Software


Many of the calculations made by the software are invisible to the user. Therefore, a
basic understanding of how the system processes the reservoir data is necessary.
Fig. 44 illustrates the components of the system. Note that the geometry models are
incorporated to simulate all aspects of the fracture mechanics. NODAL analysis and
the application of the NPV concept for economics complete the system package.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 41 of 65

Fig. 44. Components of the NPV calculation.


The reservoir is assumed to be a horizontal, isotropic, homogeneous, porous
medium bounded by the top, bottom, and outer impermeable strata. Also, the
reservoir has constant initial pressure, permeability, porosity and thickness. The
geometry of the drainage region can be a square or rectangle. A finite-conductivity
fracture contacts the well and penetrates the entire vertical extent of the formation.
This fracture is assumed to have a constant permeability, porosity, and width. The
flow entering the wellbore comes only through the fracture and obeys Darcy's law in
the entire system. The properties of the reservoir and fracture are independent of
pressure. Fig. 45 is a plan view of the physical system modeled, and shows a
fracture in a bounded reservoir.

Fig. 45. Fracture in a bounded reservoir.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 42 of 65

A method is now needed to predict the production. Remember from the previous
discussion that IPR curves are used for this purpose. The software uses constant-
rate drawdown type curves for a well located at the center of a closed square
reservoir. Fig. 46 through Fig. 48 show the type curves that exist numerically within
the software.

Fig. 46. Constant-rate type curve for finite-conductivity fracture


 closed square system (xe/ye = 1).

Fig. 47. Constant-rate type curve for finite-conductivity fracture


 closed rectangular system (xe/ye = 2).

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 43 of 65

Fig. 48. Constant-rate type curve for finite-conductivity fracture


 closed rectangular system (xe/ye = 4).
The software uses these type curves to calculate the IPR curves based on the
critical variables one inputs to the various parameter screens. Remember also that
the transient pressure behavior can last for a long period of time, especially in low-
permeability reservoirs. Notice on the type curves that since the reservoir is
bounded, with no flow across the boundary, the curves go upward. The point at
which each curve on the plot starts to bend upward is the end of the transient flow
and the beginning of the pseudosteady-state flow. Also notice that the early time
behavior is influenced primarily by the fracture conductivity, and that the late time
pressure behavior is mostly dependent on the fracture penetration and reservoir
geometry. Table 2 shows the fracture lengths for various drainage shapes noted on
the type curves.
Table 2. Maximum Fracture Lengths for Drainage Shapes
Drainage Area Square Shape
(acres)
1x2 1x4
40 660 933 1320
80 933 1320 1866
160 1320 1866 2640
320 1866 2640 3733
640 2640 3733 5280

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 44 of 65

The optimum fracture length is a function of the reservoir properties and not the
drainage area assigned to the well. Therefore, if the software indicated an optimum
length of 1000 ft, for example, and the client has used 40-acre spacing, Table 2
shows the best shape would be either 1 x 2 or preferably 1 x 4. This shows that to
fully exploit the reservoir, the wells should be spaced in a rectangular pattern. The
fracture azimuth must be known to plan this type of spacing. If these wells have
already been drilled in a square pattern, the optimum fracture length is still not
necessarily 660 ft. The azimuth of the fractures should be determined and then the
drainage boundaries reconsidered.
To optimize the production, each component must be analyzed separately and then
as a combination to evaluate the entire system. Fig. 49 shows an example IPR
curve that the software might generate internally for various fracture lengths and
Fig. 50 shows a tubing intake curve for two different pressures.

Fig. 49. Transient IPR for a fracture well in a closed square reservoir.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 45 of 65

Fig. 50. Tubing intake curve.


Using all of the input data, the analysis is made and one may choose to plot a family
of curves similar to those shown in Fig. 51 through Fig. 54. The graphics screen
shows the options available for immediate plots. Parametric studies can also
provide valuable information for evaluating the effect one variable may have on
another. Holding the NPV and fracture half-length as the axis, the effect of the
various tubing sizes, formation permeabilities, injection rates, fracturing fluid
viscosities, fracture heights, leakoff coefficient and many others can be compared
and used for redesign if necessary.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 46 of 65

Fig. 51. FracNPV analysis for one, two, and three years versus fracture length.

Fig. 52. Cumulative production versus fracture length.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 47 of 65

Fig. 53. Production decline versus time.

Fig. 54. Cumulative production versus time.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 48 of 65

There are some cautions and design considerations that need to be mentioned when
working with the software. These cautions and considerations are stated as follows.
• A single-phase simulator was used to generate the type curves used in the
software, and Vogel's correlation is applied below the bubblepoint pressure. It is
very important to use the correct pressures.
• Multiple fracturing fluids may be input, but the software will not pick the best one
for the treatment. The software uses the fluid that is most effective for the
particular segment of length. This feature is useful when trying to obtain the
highest pump rate possible, by varying the fluid viscosity, in cases where the net
pressure is a limiting factor. The software cannot simulate viscosity degradation,
and assumes the first fluid has degraded to the values of the second fluid, etc. It
may be best to use only one fluid in a rerun, that fluid being the one shown for the
length desired. The exception would be those designs where one is actually
going to use a variable fluid viscosity system.
• The best proppant is not picked if more than one is listed on the screen. Each
proppant is considered a separate case and printed out as such. NOTE: the
graphics produced are only for the first proppant listed.
• The program does not contain the two-phase material balance calculations for
the long-term prediction of the production.
• The type curves used in the program are very good for low-permeability
reservoirs (1 md or less), and depending on the actual value of the permeability
may be accurate to five years. For permeabilities over 1 md, the transient time
could be much less than the life of the well; therefore, the accuracy of the
production rate decreases. This latter case, again depending on the actual
permeability, could be accurate for possibly only one year.
The application of the FracNPV software exists whenever a propped fracture
stimulation treatment is designed. Software is constantly upgraded, so the reader
should periodically examine the applicable documentation. Also, many good
examples of treatments run using the software are available in the references and
the reader should study these as examples. The inclusion of examples such as
these is too voluminous for the purpose of this section topic.

5 Application
The primary goal of this manual section (Reservoir Evaluation) was to acquaint the
reader with the technology involved in obtaining those variables that are so critical
for designing the best fracturing treatment, and then being able to evaluate that
treatment using the same technology. Many tools and methods were shown to make
the job easier. Learn who and where to go for help in using some of the more
complex systems.
The method of reservoir evaluation, from obtaining the critical variables to evaluating
the actual treatment, might be listed and summarized as follows.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 49 of 65

1. Design and complete a valid well-performance test prior to any hydraulic


fracturing. The design should consider both an appropriate flow time and shut-in
time to obtain the needed data. Sometimes this well test may be a drill-stem test
(DST).
2. Construct a log-log grid corresponding to the grid on the type curves. Plot the
data (in the appropriate pressure and time functions) on the grid to construct the
diagnostic plot.
3. Study the log-log plot for the pattern recognition of the wellbore storage, flow
regimes, fractures, reservoir type, boundaries, etc. Use the specialized plot(s) if
necessary to help identify some of the characteristics if unsure.
4. Select the basic type curve and place the log-log plot over it. Keep the axis of the
two graphs parallel and obtain the best fit. Record the match points and
corresponding dimensionless time and pressure. Calculate the desired variables
using the type-curve equations. Check for consistency.
5. Design an optimized fracturing treatment using the FracNPV software. A very
good estimation of the gross fracture height is needed, and the fracture-height
log would be a good source. Parametric studies can be made to examine the
influence of one set of variables on another set to help choose the correct
fracturing fluids, proppants, etc.
6. Execute the treatment exactly as designed, monitoring and recording the critical
parameters. Tag the fracturing fluid or use any method that will allow the
fracture-height determination after the job. If treatment differs from the actual
design, then rerun the actual treatment through the software to obtain a new
production estimate.
7. Track the production results at least monthly. This is where most of an
evaluation falls apart. Once the treatment has been pumped, the tendency is to
forget about it and go on to the next one. The evaluation will never progress
unless an effort is made to follow through. Not only is it important to know that
the production response was very good, but also if the response was not what
one predicted. Fig. 55 is a form that might be considered for use in tracking the
well performance and for having a summary of the job parameters and critical
variables. Fig. 56 is an example of the tracking form used on an actual well.
8. If production results are not as predicted by the design or the modified design
after knowing the fracture height, then design a well performance test to evaluate
the reservoir containing a fracture. It is very helpful to have the original
permeability from an unfractured case to use to help determine the other
parameters in a fractured case. This is the reason a well test was recommended
prior to fracturing. Knowing the reservoir is now fractured, the mechanics and
duration of the actual test can be designed.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 50 of 65

9. Repeat the process as in Step 1 through Step 4, being especially watchful for the
characteristic finite conductivity of a fracture. Calculate the fracture dimensions
and determine if they are realistic. If not realistic, review the test analysis and
fracturing treatment for error. If realistic, run the parameters back through the
design software to see what variables need to be changed to match the
production as well as the parameters obtained from the well test. Be aware that if
the original fracturing treatment was not designed properly (insufficient proppant
at distance in the fracture), then the fracture may show up as being short on a
well test. The test always indicates the effective fracture length. The engineer
designing the treatment needs to be sure the fracture length will be effective.
Reservoir evaluation is a must for an effective hydraulic fracture design and for
determining if the actual fracturing treatment was successful. The treatment was
successful if it provided the production predicted by the process the user designed.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 51 of 65

Fig. 55. Well performance tracking form.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 52 of 65

Fig. 55 (back).

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 53 of 65

Fig. 56. Well performance tracking form.

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 54 of 65

Fig. 56 (back).

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 55 of 65

6 Equation Summary

6.1 Oil IPR Equations

6.1.1 Darcy's Law

7.08 × 10 −3 kh( pr − pwfs )


qo =
 r  3 
µ o Bo ln e  − + s 
  rw  4 
q q 7.08 × 10 − 3 kh
PI = = =
pr − pwfs ∆p  r  3 
µ o Bo ln e  − + s 
  rw  4 
AOF = ( PI )( pr − 0)
Where:
q = oil flow rate (B/D)
AOF = absolute open flow potential (B/D)
k = permeability (md)
h = net vertical formation thickness (ft)
pr = average formation pressure (shut-in BHP) (psi)
pwfs = average flowing bottomhole pressure at the sandface (psi)
µo = average viscosity (cp)
Bo = formation volume factor (res bbl/STB)
re = drainage radius (ft)
rw = wellbore radius (ft)
S = skin factor (dimensionless)
PI = Productivity Index (B/D/psi)

6.1.2 Vogel Test Data ( Pr ≤ pb )

2
qo pwfs  pwfs 
= 1 − 0.2 − 0.8 
qo max pr  pr 

6.1.3 Combination Vogel = Darcy Test Data ( Pr ≥ pb )

1. For test when pwftest >pb

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 56 of 65

q
PI =
Pr − pwfs
qb = PI ( pr − pb )
PI × pb
qo max = qb +
1.8
Points on IPR curve
For pwf > pb :
qo = PI ( pr − pwf )
For pwf < pb :
  pwf   pwf  
2
qo = qb + ( qo max − q b ) × 1 − 0.2   − 0.8   
  pb   pb  
 
2. For test when pwftest< pb
q
PI =
pb   pwf  
2
 pwf 
( pr − pb ) + 1 − 0.2   − 0.8   
1.8   pb   pb  
 
qb = PI ( pr − pb )
PI × pb
qo max = qb +
1.8
Points on IPR curve
For pwf > pb :
qo = PI ( pr − pwf )
For pwf < pb :
  pwf   pwf  
2
( qo max − qb ) × 1 − 0.2   − 0.8   
  pb   pb  
 
Where:
qo= flow rate (B/D)
qb= flow rate at bubblepoint (BD)
pb= bubblepoint pressure (psi)
qomax= maximum flow rate (Vogel or combination) (B/D)
PI= Productivity Index (B/D/psi)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 57 of 65

6.1.4 Jones IPR

pr − pwfs = aq 2 + b
    re    
 2.30 × 10  µ o Bo ln 0.472  r   + s  
− 14
βBo2ρ     w  
pr − pwfs =  q2 +  q
 h p2rw   7.08 × 10 − 3 kh 
   
 

− b ± b 2 + 4 a ( pr − 0 )
AOF =
2a
Where,
 2.30 × 10 −14 βB 2ρ 
a= o  2
q +
 2 
 h p rw 
    re    
 µ o Bo ln0.472   + s  
   rw    
b = 
7.08 × 10 − 3 kh q
 
 

q = flow rate (B/D)


pr = average reservoir pressure (shut-in BHP) (psi)
pwfs = flowing BHP at sandface (psi)
β = turbulence coefficient (ft-1)
2.33 × 10 10
β= ( after Katz )
k 1. 201
Bo = formation volume factor (res bbl/STB)
ρ = fluid density (lbm/ft3)
hp = perforated interval (ft)
µo = viscosity (cp)
re = drainage radius (ft)
rw = wellbore radius (ft)
S = skin factor (dimensionless)
k = permeability (md)
a = turbulence term

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 58 of 65

b = darcy flow term

6.2 Gas IPR Equations

6.2.1 Darcy's Law (Gas)

703 × 10 −6 kh( pr2 − pwfs


2
q=
 r  3 
µ T Z ln  e  − + s 
  rw  4 
Where:
q = flow rate (Mcf/D)
k = permeability (md)
h = net vertical thickness (ft)
pr = average formation pressure (shut-in BHP) (psia)
pwfs = sandface flowing BHP (psia)
µ = viscosity (cp)
T = temperature (°R)
Z = supercompressibility (dimensionless)
re = drainage radius (ft)
rw = wellbore radius (ft)
S = skin factor (dimensionless)

6.2.2 Jones' Gas IPR (General Form)

pr2 − pwfs
2
= aq 2 + bq
  r  
− 12 1.424 × 10 3 µ TZ ln 0.472 e  + s 
3.16 × 10 βγ g TZ 2   rw  
pr2 − pwfs
2
= 2
q + q
h p rw kh

− b ± b 2 + 4a( pr2 )
AOFP =
2a

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 59 of 65

Where:
3.16 × 10 −12 βγ g TZ
a=
h p2 rw
  r  
1.424 × 10 3 µ TZ ln 0.472 e  + s 
b=   rw  
kh
q = flow rate (Mcf/D)
a = turbulence term
b = darcy Flow term
p r = reservoir pressure (shut-in BHP) (psia)
pwfs = sandface flowing BHP (psia)
β = turbulence coefficient (ft-1)

2.33 × 10 10
β=
k 1. 201
γg= gas specific gravity (dimensionless)
T = reservoir temperature (°R)
hp= perforated interval (ft)
µ = viscosity (cp)
re = drainage radius (ft)
rw = wellbore radius (ft)

6.3 Backpressure Equation

( )
n
q g = c pr2 − pwfs

Where:
703 × 10 −6 kh
c=
 r  3 
µ TZ ln  e  − + s 
  rw  4 
n = 0.5
qg = flow rate (Mcf/D)
k = permeability (md)
h = net vertical thickness (ft)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 60 of 65

p r = average formation pressure (shut-in BHP) (psia)


pwfs = sandface flowing BHP (psia)
µ = viscosity (cp)
T = temperature (°R)
Z = supercompressibility (dimensionless)
re = drainage radius (ft)
rw = wellbore radius (ft)
S = skin factor (dimensionless)

6.4 Transient Period Equations

6.4.1 Time to Pseudosteady State

 φµct re2 
tstab = 948 
 k 
Where:
φ = porosity (fraction)
µ = viscosity (cp)
ct = total system compressibility (psi-1)
re = drainage radius (ft)
k = permeability (md)
tstab = time for pressure transient to reach re (hr)

6.4.2 Oil IPR (Transient)

qo =
(
kh pr − pwfs )
  kt  
162.6 µ o Bo log 2
 − 3. 23 + 0.87 S 
  φµct rw  
Where:
k = permeability (md)
h = net vertical thickness (ft)
µ = viscosity (cp)
Bo = formation volume factor (res bbl/STB)
t = time of interest; t<tstab (hr)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 61 of 65

φ = porosity (fraction)
ct = total system compressibility (psi-1)
rw = wellbore radius (ft)
S = skin factor (dimensionless)

6.4.3 Gas IPR (Transient)

qg =
(
kh pr2 − pwfs
2
)
  kt  
1,638 µ TZ log  2
 − 3. 23 + 0.87 S 
  φµct rw  
Where:
qg = flow rate (Mcf/D)
k = permeability (md)
p r = reservoir pressure (shut-in BHP) (psia)
pwf = flowing bottomhole pressure at sandface (psia)
µ = viscosity (cp)
T = temperature (°R)
Z = supercompressibility (dimensionless)
t = time of interest; t<tstab (hr)
f = porosity (fraction)
ct = total system compressibility (psi-1)
rw = wellbore radius (ft)
S = skin factor (dimensionless)

6.5 Completion Pressure Drop Equations

6.5.1 Gravel-Packed Wells


1. Oil Wells (General)
pwfs − pwf = ∆p = aq 2 + bq
9.08 × 10 − 13 β Bo2 ρL µ Bo L
∆p = q2 + q
A2 1.127 × 10 − 3 k g A

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 62 of 65

Where:
9.08 × 10 −13 β Bo2 ρ L
a=
A2
µ Bo L
b=
1.127 × 10 − 3 k g A

q = flow rate (B/D)


pwf = pressure well flowing (wellbore) (psi)
pwfs = flowing BHP at sandface (psi)
b = turbulence coefficient (ft-1)
For GP Wells:
1.47 × 10 7
β=
k 0.55
g

Bo = formation volume factor (res bbl/STB)


ρ = fluid density (lbm/ft )
3

L = length of linear flow path (ft)


A = total area open to flow (ft2)
(A = area of one perforation x shot density x perforated interval),
kg = permeability of gravel (md)
2. Gas Wells (General)

pwfs2 − pwf2 = ∆p = aq 2 + bq
1.247 × 10 − 10 β γ g TZL 8.93 × 10 3 µ TZL
pwfs2 − pwf2 = q2 + q
A2 kg A

Where:
1.247 × 10 −10 βγ g TZL
a=
A2
8.93 × 10 3 µ TZL
b=
kg A

q = flow rate (Mcf/D)


pwfs = flowing pressure at the sandface (psia)
pwf = flowing bottomhole pressure in wellbore (psia)
β = turbulence factor (ft-1)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 63 of 65

1.47 × 10 7
β=
k 0.55
g

γg = gas specific gravity (dimensionless)


T = temperature (°R)
Z = supercompressibility (dimensionless)
L = linear flow path (ft)
A = total area open to flow (ft2)
(A = area of one perforation x shot density x perforated interval),
µ= viscosity (cp)

6.5.2 Open Perforation Pressure Drop


1. Oil Wells (General)
p wfs − pwf = ∆p = aq 2 + bq
 − 14 2  1 1    rc  
 2.30 × 10 β Bo ρ r − r    µ Bo  ln r  
 p c 2  p 
∆p =  q + q
 2
Lp   7.08 × 10 − 3 L p k p 
   
   
Where:
 1 1
2.30 × 10 − 14 β Bo2 ρ − 
 rp rc 
a=
L2p

 r 
µ Bo  In c 
 rp 
b=
7.08 × 10 − 3 L p k p

qo = flow rate/perforation (q/perforation) (B/D)


β = turbulence factor (ft-1)
2.33 × 10 10
β=
k 1p.201

Bo = formation volume factor (res bbl/STB)


ρ = fluid density (lbm/ft )
3

Lp = perforation tunnel length (ft)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Reservoir Evaluation Dowell
Page 64 of 65

µ = viscosity (cp)
kp = permeability of compacted zone (md)
kp = 0.1 k formation if shot overbalanced,
kp = 0.4 k formation if shot underbalanced,
rp = radius of perforation tunnel (ft)
rc = radius of compact zone (ft)
(rc = rp + 0.5 in.).
2. Gas Wells (General)

2
pwfs − pwf
2
= ∆p = aq 2 + bq
 −12  1 1   rc  
 3.16 × 10 β γ g TZ  r − r    1.424 × 10 µ TZ  ln r  
3
 p c 2   p 
= q + q
 L2p   k p Lp 
   
   
Where:
 1 1
3.16 × 10 − 12 β γ g TZ  − 
 rp rc 
a=
L2p

 r 
1.424 × 10 3 µ TZ  In c 
 rp 
b=
Lpk p

qo = flow rate/perforation (q/perforation) (B/D)


β = turbulence factor (ft-1)
2.33 × 10 10
β=
k 1p.201

γg = gas specific gravity (dimensionless)


T = temperature (°R)
Z = supercompressibility factor (dimensionless)

DOWELL CONFIDENTIAL
Section 100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Reservoir Evaluation
Page 65 of 65

rc = radius of compact zone (ft)


(rc = rp + 0.5 in.),
rp = radius of perforation (ft)
Lp = perforation tunnel length (ft)
µ = viscosity (cp)
kp = permeability of compacted zone (md)
kp = 0.1 k formation if shot overbalanced,
kp = 0.4 k formation if shot underbalanced.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 1 of 52

ROCK MECHANICS

1 Introductory Summary............................................................................................................. 3

2 Lithology................................................................................................................................... 3

3 Basic Concepts ........................................................................................................................ 9


3.1 Stress Concept .................................................................................................................... 9
3.2 Stress Definition ................................................................................................................. 10
3.3 Strain.................................................................................................................................. 17
3.4 Modulus/Poisson Effect...................................................................................................... 18
3.5 Effective Stress .................................................................................................................. 22
3.6 Failure Criteria ................................................................................................................... 23

4 Measurement of Rock Properties ......................................................................................... 27


4.1 Typical Properties .............................................................................................................. 29
4.2 Toughness ......................................................................................................................... 31

5 Determining In-Situ Stress .................................................................................................... 35


5.1 Core Tests ......................................................................................................................... 35
5.2 Microfracturing ................................................................................................................... 42
5.3 Pump-In/Flowback Test ..................................................................................................... 44
5.4 Logs ................................................................................................................................... 44

6 Fracture Height (Post) .......................................................................................................... 47


6.1 Radioactive Tracer............................................................................................................. 47
6.2 Temperature ...................................................................................................................... 50

FIGURES

Fig. 1. Example Litho-Density log and CNL* Compensated Neutron log..................................... 6


Fig. 2. Example crossplot chart.................................................................................................... 7
Fig. 3. Lithology log...................................................................................................................... 8
Fig. 4. Triaxial stress.................................................................................................................. 10
Fig. 5. Randomly oriented plane. ............................................................................................... 11
Fig. 6. Rectangular axis system. ................................................................................................ 12

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 2 of 52

Fig. 7. Two-dimensional systems. ..............................................................................................13


Fig. 8. Stress charts....................................................................................................................15
Fig. 9. Strain measurement. .......................................................................................................17
Fig. 10. Typical stress/strain curves. ..........................................................................................18
Fig. 11. Stress strain relationship. ..............................................................................................20
Fig. 12. Compressional force......................................................................................................21
Fig. 13. Principal stresses in a triaxial application. .....................................................................23
Fig. 14. Mohr's circle. .................................................................................................................24
Fig. 15. Mohr failure envelope. ...................................................................................................25
Fig. 16. Typical failure envelope.................................................................................................26
Fig. 17. Dynamic elastic properties. ...........................................................................................27
Fig. 18. Permeability versus confining stress for a fractured rock specimen..............................30
Fig. 19. Stress concentration near the tip of a crack. .................................................................32
Fig. 20. Modified ring. .................................................................................................................33
Fig. 21. Load/displacement curve obtained during a Modified Ring Test...................................33
Fig. 22. Drawing of the LVDT* displacement gauge developed at the Sandia National
Laboratory1....................................................................................................................36
Fig. 23. Typical ASR curve. ........................................................................................................37
Fig. 24. Anelastic strain recovery measurements.......................................................................38
Fig. 25. Gage pattern. ................................................................................................................39
Fig. 26. Typical DSCA plots........................................................................................................40
Fig. 27. Stereoplot. .....................................................................................................................41
Fig. 28. Idealized plot. ................................................................................................................42
Fig. 29. Microfrac test. ................................................................................................................43
Fig. 30. Fracture-height log. .......................................................................................................46
Fig. 31. Two-isotope tracking of a two-stage fracture treatment. ...............................................48
Fig. 32. Two-isotope tracking of a single-stage fracture treatment.............................................49
Fig. 33. Temperature survey with RA tracer...............................................................................51

TABLES

Table 1. X-Ray Diffraction Analysis ..............................................................................................4


Table 2. Dynamic Versus Static Tests........................................................................................28

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 3 of 52

1 Introductory Summary
Rock mechanics, as applied to petroleum engineering, has primarily been focused
on explaining the orientation of fractures (Hubbert and Willis, 1957). It has only been
recently that the importance of rock mechanics in several areas of the oil and gas
industry was understood. The mechanical behavior influences every aspect of
completion, stimulation and production. The work on fracture orientation was
significant because it indicated the importance of differential stresses within the
earth's crust. These stresses affect the behavior of rock in different ways depending
on the type of disturbance associated with the area, as well as any man-made
disturbances.
The National Academy of Sciences defines rock mechanics as “the theoretical and
applied science of the mechanical behavior of rock; it is that branch of mechanics
concerned with the response of the rock to the force fields of its physical
environment.” This definition applies to mining and many aspects of civil engineering
as well as the application in petroleum engineering, and specifically to the
optimization of hydraulic fracturing. As deeper completions were attempted, collapse
of the borehole and evidence of other problems became more common. The causes
of these problems were instabilities due to large tectonic forces. Rocks are not inert
and have a definite behavior that can be influenced by man-made disturbances
associated with all aspects of drilling and production.
A discussion covering rock mechanics must have a broad scope not only because of
the various engineering applications, but also because of the wide variety of material
types that are classified as “rock.” Some rocks can display a brittle, elastic nature at
confining pressures of thousands of psi, while other rocks become plastic at a much
lower pressure. Other rocks such as friable sandstones and certain shales are
weakened by the presence of water. Rock salt and gypsum respond plastically at
low confining pressures and are also highly soluble. These differences are only a
small part of the large body of technology surrounding the application of rock
mechanics to the petroleum industry. Therefore, this section will deal more with the
concepts needed for the reader to have a better understanding of stresses and the
distribution of these stresses, how rocks fracture, how certain test data are obtained,
and how data can be used to improve fracture design.
2 Lithology
The individual character of a rock in terms of mineral composition, structure, and
texture is termed the lithology. Rocks vary greatly in general appearance and also in
the detail of their composition. Even clay, for example, may appear to be a formless
mass but is actually a maze of particles with a regular form. Sandstone is composed
of some form of silica and may contain some of the clay mineral kaolinite, or other
minerals. Therefore, rocks are said to be composed of minerals. Commonly,
minerals are in the form of perfect crystals composed of elements such as silicon,
oxygen and aluminum. At other times no crystalline structure can be observed
(amorphous). These elements and the minerals that determine lithology can be

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 4 of 52

identified in the laboratory using various analytical techniques. Table 1 shows an


example analysis on several feet of core using X-ray diffraction. The clay weight
percent is for total clays, and can be further analyzed to obtain the specific clays
such as illite, chlorite, and kaolinite.
Table 1. X-Ray Diffraction Analysis
Depth (ft) Clays Quartz Plagioclase Calcite Siderite Pyrite
“Uppermost White” Sandstone
13567.3 97 2 1
“Uppermost Reservoir Facies” Sandstones
13567.3 15 75 10
13568.7 17 62 3 18
13570.4 13 84 3
13571.8 15 80 4 1
13574.1 18 77 5
13575.0 15 76 9
13585.3 15 77 8
13586.2 14 79 7
13591.8 17 79 4
_
X= 15.4 76.6 5.9 0.1 1.8
“Lower Reservoir Facies” Sandstones
13594.5 15 82 3
13595.6 14 83 3
13599.6 16 81 3
13600.1 97 3
_
X= 11.3 85.8 3.0
Interbedded Shales
13550.3 74 22 4
13555.6 35 47 9 9
13560.5 45 51 4
13565.9 60 29 6 5
13579.9 45 48 7
13589.1 64 25 7 4
13600.3 64 20 6 7 3
13601.6 53 33 4 10
13602.4 57 34 4 5
_
X = 55.2 34.3 5.7 4.4 0.3
_
X = Average rock composition.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 5 of 52

Rocks are also grouped according to origin such as igneous (molten at one time),
metamorphic (transformed by heat and/or pressure, that is, limestone to marble) and
sedimentary (deposited in layers). Sedimentary rocks are by far the group that is
most interesting to the petroleum industry. The sedimentary rocks represent large
thick layers accumulated over a vast period of time, and the processes that formed
these layers are still active on the earth today.
There are also two major divisions of sedimentary rocks  (1) the clastic rocks that
are composed of particles derived from erosion and weathering of older rocks, and
(2) the chemical and organic rocks that are composed of materials that have either
been precipitated chemically or accumulated by the activities of plants and animals.
Examples of common clastic rocks would be conglomerate, sandstone (including
arkose), and shale. The chemical and organic rocks would be limestone, dolomite,
salt, gypsum, coal, and diatomites. Rocks also differ in the way the minerals or
mineral grains are bound together in the rock mass. A very stable cementing
material is silica, and many sandstones and conglomerates have this as the binding
agent. The soluble or slightly soluble systems would be those sandstones or
conglomerates that are cemented with calcite or gypsum. Friable sandstones are
incompletely or weakly cemented, and the totally uncemented rocks are usually the
clay-bound sandstones. The very soft soil-like rocks (such as compaction shale,
chalk and marl) may also be included as examples of poor cementation.
Obviously, the type of rock and the way it is cemented will make a difference in an
analysis. The differences can also impact the strength of the rock as can the
porosity and related permeability. The degree and type of fluid filling the pore
spaces will also make a difference in test results and will be discussed later. All of
these related characteristics are important when screening data for use in the design
of hydraulic fracturing. The entire set of data to describe lithology can best be
obtained from a representative core taken from the formation of interest. This core
should be oriented whenever possible for other reasons also discussed later. Logs
would be the next best method to obtain the lithology. Which method is chosen will
depend on the need to have exact versus approximate data. Logs can be used to
determine the type of rock (sandstone, limestone or shale) but not the exact mineral
composition of the rock as shown in Table 1. Logging technique and all of the
interpretive methods used to obtain the lithology are beyond the scope of this
section, and if interested the reader should study the interpretation principles and
charts that are available. Therefore, only a few methods will be discussed here.
One method to obtain the lithology is to use the combined neutron formation density
log. Fig. 1 shows an example Litho-Density* Log and CNL* Compensated Neutron
Log with gamma rays and another trace labeled pef (photoelectric factor). The Litho-
Density Log is the trace labeled “density porosity” and is obtained by emitting
gamma rays into the formation. This determines an electron density which can be
related to the bulk density and porosity. The CNL is the trace labeled “neutron
porosity” and is obtained basically by the response of the tool while emitting high-

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 6 of 52

energy neutrons into the formation. Lithology is determined by taking the value of
the neutron porosity and the density porosity at the point of interest on the log and
entering these data on a crossplot. An example crossplot is given in Fig. 2. The
crossplot chart is entered with the point porosities computed as if the matrix had the
same properties as a water-saturated limestone; as a result, the limestone is a
straight line of equal density and porosity. Mixed lithology points will fall between the
corresponding lithology lines. Points falling between the lines can be volumetrically
proportioned to obtain the percent of each material. These charts were constructed
for boreholes filled with water or water-base mud and for clean liquid-saturated
formations.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 7 of 52

Fig. 1. Example Litho-Density∗ log and CNL* Compensated Neutron log.

Fig. 2. Example crossplot chart.


Another method is to use the pef trace from Fig. 1 and to enter another type of
crossplot where the sandstone, dolomite and limestone lines are vertical on the plot.
Density is the Y axis and the photoelectric factor is the X axis. This is similar to the
other crossplot but may be a little easier to use.
The final method to be discussed here is the computer-generated lithology log
shown in Fig. 3. Essentially, the computer makes all of the calculations discussed in
the previous methods, and presents the data by rock type as well as providing traces
for the gamma ray, resistivity and porosity. Hydrocarbon and water are also

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 8 of 52

indicated on the log. Obviously, because work by hand to crossplot data is not
necessary and the margin for human error is less, this would be the log of choice for
someone unfamiliar with the total computational process.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 9 of 52

Fig. 3. Lithology log.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 10 of 52

3 Basic Concepts
The hydraulic fracturing technique was introduced in 1948 and has been one of the
major developments in the petroleum industry. Prior to this time, however, pressure
parting in water injection wells, lost circulation while drilling, and formation
breakdown while squeeze cementing had all been observed and documented. A
popular theory was that the pressure applied at the wellbore had parted the
formation along a bedding plane. This implies that the entire overburden must be
lifted, although in the cases where the pressure was known it was significantly less
than the overburden pressure determined by density. Later it was pointed out
(Hubbert, 1953) that the normal state of underground stress is one of unequal
principal stresses; in tectonically relaxed areas, characterized by normal faults, the
least stress should be horizontal. This means that in most cases fracturing should
be possible with pressure less than the overburden, and that these fractures should
be vertical.

3.1 Stress Concept


The early observations of fracturing previously mentioned assumed that the stress
field was nearly hydrostatic. This means that the principal stresses were assumed to
be nearly equal, with the value of the stresses being that of the overburden pressure.
This assumption was not uncommon in early attempts to understand the mechanical
behavior of rock. However, rocks are not ideally homogeneous, continuous,
isotropic (nondirectional in properties), linear, and elastic. The earth has shown
considerable movement over the long periods of geologic time, and the rocks have
repeatedly been deformed to the limit of failure by folding and faulting. Substantial
differences in the principal stresses are required for this folding and faulting to take
place. Therefore, the in-situ general stress condition is one in which there are three
unequal perpendicular principal stresses. It is necessary to understand the stress
conditions to provide a basis for determining the fracture behavior, and ultimately for
treatment optimization.
Increasing fluid pressure applied to a local stress field in which the principal
perpendicular stresses are unequal will eventually cause rupture or parting of the
rock. The plane along which this fracturing is first possible is the one perpendicular
to the least principal stress (or parallel to the plane of greatest stress). Fig. 4
illustrates the concept of triaxial stresses and the preference of the fracture plane.
The in-situ stress field is the most important parameter controlling the fracture
orientation as well as the fracture-height growth versus containment.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 11 of 52

Fig. 4. Triaxial stress.

3.2 Stress Definition


Stress analysis is really just a matter of pure statics and is independent of the plastic,
elastic, viscous or other type of properties assumed for the material. The Greek
letter “sigma” (σ) is used to designate the components of normal stress (σx, σy and σz)
as well as the principal stresses (σ1, σ2 and σ3). A component of stress that acts
perpendicular (normal) to a plane is designated as the normal stress. A component
of stress also acts parallel to a plane and represents the forces tending to slide or
shear the material in the plane. This stress component is called shear stress and is
designated by the Greek letter “tau” (τ).
Fig. 5 shows a randomly oriented plane of area designated as δA (delta A) and a
point O in the interior. The forces acting at point O are described as follows. For
each direction of OP through point O, assume that the material is cut across a small
area of δA in a plane through O and that this plane is normal to OP. The surface of
the cut on the side of P is called the positive side, and that on the side opposite of P
is called the negative side. The effect of internal forces existing in the body across
the surface δA is equal to a force, δF, exerted by the material on the positive side of
the surface on the negative side (and an equal, oppositely directed force exerted by
the material on the negative side on the positive side). Therefore, this stress
quantity is expressed as a force per unit area. The limiting value of the ratio δF/δA
as δA approaches zero is the stress vector at the point O across the plane that is
normal in the direction OP. This vector pOP is given as follows.

lim δF
POP = δA → 0 (1)
δA

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 12 of 52

There is a convention of the sign when forces are designated. Forces are termed
positive when compressive (direction shown by δF in Fig. 5). This is opposite to the
convention used in working with the theory of elasticity where stresses are
designated as positive when tensile. The positive sign for compressive stresses is
more convenient, because stresses caused by the depth of burial, confining pressure
in the laboratory test equipment and pore pressure are all compressive.

Fig. 5. Randomly oriented plane.


Looking at Fig. 5 again, it should be obvious that an infinite number of planes can be
drawn through point O. Although the resultant force acting on these planes is the
same, the stresses acting on the individual planes are different because of the
various inclinations. Also, for a complete description of the state of stress it is
necessary to specify the magnitude, direction, sense and surface on which the stress
acts. Consider, which is a right-handed system of rectangular axes Ox, Oy and Oz at
O. Take OP in the direction of Ox and the vector pox will have components in the x, y,
and z directions which can be written as σx, τxy and τxz. Because the area (δA) is
perpendicular to Ox, the stress δx is called a normal stress. Note that τxy and τxz are
both in the plane of area δA and are the shear stresses that tend to shear or slide the
material in the plane δA. Repeating this for the other directions (y and z) also
produces one normal stress and two shear stresses. The total nine quantities
produced in this manner are called the stress components at point O. However, it
can be shown that τxy = τyx, τyz = τzy and τzx = τxz, which actually reduces the stress
components to six. The stress vector pOP for any direction of OP can now be
expressed in terms of these six stress components.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 13 of 52

Fig. 6. Rectangular axis system.


Frequently, the analysis of stress is done in two dimensions. Stress in two
dimensions is discussed here, because it is relatively easy and it is hoped it will
clarify some of the concepts discussed earlier. The two-dimensional system uses
only the xy plane, so everything is independent of z. Considering the xy plane as the
plane of the paper, Fig. 7a shows the stress components σx and τxy and Fig. 7b
shows the stress components σy and τyx. The stress component relationship and the
variation of the stress with direction can now be examined.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 14 of 52

Fig. 7. Two-dimensional systems.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 15 of 52

Consider an area of material that is at rest and in equilibrium. The forces exerted by
the stresses over the surface of this area must also be in equilibrium. Fig. 8a is a
square (OABC) with a very small side length of a, and shows the forces per unit area
that are exerted on the sides of the square by the material beyond it. It may seem
from examination of the figure that the force on the square in the x and y directions is
zero, but a couple exists per unit length perpendicular to the plane of the paper and
exerted on the square in the plane of the paper.
a (τxy - τyx)
2
(2)
As mentioned previously, τxy = τyx and the condition of the equilibrium of forces is
met. The effect of this condition of equilibrium is to reduce the four stress
components, in the two dimensions, to only three (the notations for these stresses
continue in many cases simply to preserve symmetry in equations).
Fig. 8b is now used to develop the equations for calculating the normal and shear
stress on a plane. To do this, determine the components (px, py) of the stress vector
pOP in the direction OP inclined at an angle θ to Ox. This is done by considering the
equilibrium of a triangular section (perpendicular to the plane of the paper) OAB, with
the sides of the triangle given as AB = a, OB = a cos θ and OA = a sin θ. Resolving
parallel to Ox gives
AB × p x = OB × σ x + OA × t yx (3)

Canceling a factor (a),


Px = σ x cos θ + τ yx sin θ (4)

Repeating this parallel to Oy,


Py = σ y sin θ + τ xy cos θ (5)

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 16 of 52

Fig. 8. Stress charts.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 17 of 52

These are expressed in terms of the normal and shear stress (σ and τ) on the plane
AB. The positive directions are used to be consistent with those for the angle equal
to zero as in Fig. 7a. Eq. 6 can now be derived.
σ = p x cos θ + p y sin θ (6)

or
σ = σ x cos 2 θ + 2τ xy sin θ cos θ + σ y sin 2 θ (7)

that is,

( ) ( )
σ = 1/ 2 σ x + σ y + 1/ 2 σ x − σ y cos 2 θ + τ xy sin 2 θ

Also,
τ = py cos θ − px sin θ (8)

so

( )
τ = 1/ 2 σ y − σ x sin 2 θ + τ xy cos 2 θ (9)

There are always two orthogonal orientations of δA that cause the shear stress
components to vanish. They are called the principal planes. The stresses normal to
these planes are called the principal stresses, and can be calculated (in two
dimensions) by setting τ = 0 in Eq. 9. The expression for the principal stresses (σ1
and σ2) therefore becomes
1/ 2
( ) (
σ 1, 2 = 1/ 2 σ x + σ y ±  τ 2xy + 1/ 4 σ x − σ y 

2
) (10)

acting on a plane oriented at an angle given by


 σ − σx 
θ = arctan 1  (11)
 τ xy 

Eq. 10 uses the positive sign to calculate σ1 and the negative sign to calculate σ2.
The convention is that σ1 is called the major principal stress, and σ2 is the minor
principal stress.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 18 of 52

3.3 Strain
The relative position of points within a body will be altered when the body is
subjected to a stress field (because the body will deform). The very concept of
continuum mechanics is the displacement of all points (particles) of the body. The
initial position (x, y, and z) of each point is assumed to be known, and the forces
then applied to the system cause displacement to a final position. Since the sign for
displacement should follow that for stress, then displacement is positive to
correspond to positive stresses. The final objective is to determine the initial
displacement of every point from the stress and boundary conditions. Intermediate
quantities called strains are introduced to accomplish this determination. A great
deal of the theory of elasticity is concerned with strain rather than displacement, but
frequently the displacement becomes more important because it is usually what is
observed in experiments or practical applications.
Translation as a rigid body is one simple form of displacement in which the relative
positions of the points are not altered. Rotation as a rigid body around a fixed axis
is another form of displacement. If the positions of the points within a body are
changed such that the initial and final positions will not correspond by movement as
a rigid body, then the body is considered to be strained and measuring this strain
becomes necessary. Fig. 9 illustrates the two most convenient methods to measure
strain. One method illustrated is change in length, and the other is change in angle.

Fig. 9. Strain measurement.


Therefore, if l is the distance between two points (O and P) in the unstrained state
and l’ is the distance in the strained state, then
l − l′
∈= (12)
l
and is defined as the elongation at O in the direction OP. The other method is the
change in angle between two perpendicular directions (OP, OQ) and is used to

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 19 of 52

calculate strain at point O. Now, considering the strained state of the angle P′O′Q′ is
ψ′, then
γ = tan ( ψ ′ - ψ ) (13)

and is termed the shear strain. Because stresses were taken as positive in
compression, this positive shear strain reflects an increase in angle and the positive
longitudinal strain (∈) corresponds to a decrease in length.

3.4 Modulus/Poisson Effect


A rock specimen, such as a cylinder whose length is two to three times its diameter,
will deform when subjected to axial compression. The higher the stress level, the
more strain the rock will experience. The axial and lateral strains for any stress
applied can be measured by strain gages attached to the specimen. Plotting stress
against strain will produce a stress/strain curve similar to that shown in Fig. 10.

Fig. 10. Typical stress/strain curves.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 20 of 52

This curve has distinct regions, and in the first nonlinear region (initial stages of
loading from O to A ) the rock stiffens because pre-existing microcracks are closing
and minerals are slightly compressed. If the load were removed, most of the
microcracks may stay closed so that a net deformation results. The porosity of the
cracks is related to this deformation. Also, it is this region of the curve that has the
signature of the stress history undergone by the specimen during past geologic
times. This characteristic can be applied to the determination of in-situ stresses, and
will be discussed later.
Further loading (A to B) after most of the cracks have closed produces bulk rock
compression where pores deform and grains compress at a more linear rate. The
linear form is represented by the coefficient of proportionality, E, which is termed the
Young's modulus and is shown by

σ=E∈ (14)

Most rocks exhibit this response over a substantial loading range and, therefore,
Young's modulus is really a measure of the stiffness of the rock, or the parameter
expressing the resistance to deformation that a rock has for a given load condition.
Young's modulus is important because it directly affects hydraulic fracture geometry
and governs how wide the fracture will open at a given downhole pressure. The
length and width of a man-made fracture depend on the stiffness of the rock.
Continued loading beyond point B causes damage that is not reversible because
large deformations occur and the bulk modulus becomes progressively higher. The
rock behavior using a single elastic constant now becomes impossible to describe.
However, the use of a secant as well as a tangent modulus can be used to make the
problem workable. The difference in these two moduli can be significant and care
should be taken when using reported results for application to the fracture design.
The tangent Young's modulus at the in-situ stress conditions should be used for
stimulation applications, and is given as
 dσ 
Et =   (15)
 d ∈
Clarification of the stress/strain relationship is made easier by looking at portions of
the curve as shown in Fig. 11 (a, b, and c).

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 21 of 52

Fig. 11. Stress strain relationship.


The stress/strain curve (Fig. 11a) for most rocks is approximately linear in form but
ends dramatically in failure at F. It was mentioned earlier that the linear relationship
is represented by E (Young's modulus) in the form σ = E∈. However, this
representation is true only if the material is linearly elastic; Fig. 10 shows that the
curve actually has several regions prior to failure. Fig. 11b shows a perfectly elastic
material. This is defined as a unique relation, σ = f(∈), between stress and strain
that does not need to be linear. Perfect elasticity is when the material is gradually
stressed and when unloaded, the same path given by σ = f(∈) will be followed. Also,
all of the energy stored in the material while loading is dissipated during unloading.
Therefore, there is no unique modulus, but for any value of σ corresponding to a
point P, the slope PQ of the tangent to the curve is the tangent Young's modulus and
the slope of the secant OP, (σ/∈), is called the secant modulus. Fig. 11c shows the
behavior of a material that would be termed elastic. A material would be called
elastic if after loading and unloading (but prior to failure) the strain would go back to
zero but not necessarily by the same path as the loading. This effect is called
hysteresis and may be caused by energy dissipation by irreversible processes such
as the formation of new microcracks.
The beginning of this topic started with the statement that the compressive force on
a cylinder of rock will cause deformation. If the rock is compressed in one direction, it
will shorten along the direction of applied compression and expand in the lateral
directions. Fig. 12 illustrates the change in diameter with change in length due to the
compressional force.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 22 of 52

Fig. 12. Compressional force.


Compression is positive by earlier definition so in this case the lateral change, or
principal strain (∈2), must carry a negative sign because the change is one of
expansion. The stress applied to cause the compression in this case will be called
σ1 and is positive; therefore, the other principal strain (∈1) is contraction and has a
positive sign. The ratio of this lateral expansion to the longitudinal contraction is
termed Poisson's ratio and is expressed as
∈2
v= (16)
∈1

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 23 of 52

3.5 Effective Stress


A portion of the total applied stress, particularly in reservoir rock, is supported by the
fluid contained in the pores. Since the fluid is supporting some of the stress, then
the rock matrix is only affected by the effective stress component. In an actual
reservoir, fluid movement takes place and causes an effect on the mechanical
response. Diffusion of the pore fluid will also alter the results. There can be a big
difference in the results if the tests are made in an undrained (saturated) versus a
drained (depleted) condition. The speed at which loading is applied to a porous
sample can also make the sample behave either in a stiff or soft manner.
The effective stress concept was first introduced by Terzaghi in 1923 when he
suggested that a linear relationship existed between the void ratio of the material and
the logarithm of the applied pressure. Terzaghi's effective stress law for rocks was
based on saturated soils but has proved to be sufficiently accurate for engineering
purposes by Skempton in 1960. The effective stress law states that a pressure (p) in
the pore fluid of a rock will cause the same reduction in the peak normal stress
(stress required to cause failure) just as if it were caused by a reduction in a
confining pressure equal to p. The effective stress is therefore expressed as

σ′ = σ − ρ (17)

where σ is the total applied stress.


The effective stress required modification to account for the coupled
diffusion/deformation process, and the fact that the cementation between grains may
prevent the full magnitude of the pore pressure from counteracting the applied load.
The modification is made to the pore pressure and is termed the poroelastic
constant, α, and is a parameter which describes the efficiency of the fluid pressure in
counteracting the applied load. The expression with modification is

σ ′ = σ − aρ (18)

and governs the deformation of the porous medium, while the failure is controlled by
the effective stress given in Eq. 17. Derivation of the poroelastic constant is
discussed in the literature and references if the reader is interested. Typically, the
value of α for petroleum reservoirs is approximately 0.7.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 24 of 52

3.6 Failure Criteria


There is a relation between the principal effective stresses that represent a limit
beyond which instability or failure occurs. The peak stress of rock that is under
loading will increase if the rock is confined. The variation of the peak stress (σ1) with
the confining pressure (σ3) is known as a failure criterion. There are several types of
criteria discussed in the literature, and they are used for a variety of applications.
Only two of the criteria will be discussed and these are the Coulomb, or Coulomb-
Navier, and the Mohr.
Coulomb (1773) introduced one of the simplest and most important criteria. His
research into friction suggested that in the shear failure of rocks, the shear stress
causing the failure across a plane is resisted by the cohesion of the material and by
a constant times the normal stress across the plane. This criterion for shear failure
(τ) in a plane is given as τ = So + µσ where σ and τ are the normal and shear
stresses across the plane, So is a constant which is the inherent shear strength or
cohesion, and µ is another constant called the coefficient of the internal friction.
Navier modified the Coulomb criterion by letting the normal stress acting across the
plane of failure increase the shear resistance of the material by an amount
proportional to the value of the normal stress. Fig. 13 is a two-dimensional case
showing the principal stresses applied to a triaxial compressive test sample, and
should help clarify these concepts of failure.

Fig. 13. Principal stresses in a triaxial application.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 25 of 52

Letting σθ and τθ be the normal and shear stresses acting on the failure plane (AB),
the criterion says that failure will occur when the shear stress reaches

τ θ = τ t + µσθ (19)

where τt is the shear strength of the material. The angle α in Fig. 13 is the angle that
the vertical force makes with the plane of failure.
Mohr (1900) proposed that when shear failure takes place across a plane, the
normal stress and the shear stress across the plane are a functionally related
characteristic of the material. Mohr's circle diagram is probably the most important
graphical method of representing the variation of stress in two dimensions. Fig. 14
shows a Mohr circle.

Fig. 14. Mohr's circle.


The first step was to draw Cartesian coordinates where the y axis = τ and the x axis
= σ and O is the intersection. The convention is to designate the major,
intermediate, and minor principal stress by σ1, σ2 and σ3, respectively. Failure of the
rock will take place at some critical relationship between σ1 and σ3. However, since
this is two-dimensional, σ3 is not used and σ1 > σ2. Next, the value of σ1 is marked off
as OP and the value of σ2 is marked off as OQ. A circle is then drawn on diameter
PQ, where the center, C, of the circle is given as σ = ½(σ1 - σ2). Therefore, the
intersection of this circle with the horizontal axis gives the maximum and minimum
values of the normal stress. The apex gives the maximum value of the shear stress.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 26 of 52

The way the Mohr circle is most used is actually by drawing several circles using the
data obtained from a series of triaxial tests. The triaxial tests are performed under
different confining conditions until failure occurs, and a characteristic failure envelope
for a particular rock type can then be obtained. It was stated earlier that the normal
and shear stress across a plane are functionally related such as τ = f(σ). Fig. 15
shows how this relation is represented as a curve (AB).

Fig. 15. Mohr failure envelope.


However, it must be pointed out that there is no explicit formula defining this curve.
The curve is experimentally obtained by drawing it tangent to the family of circles,
producing the Mohr envelope. Failure will not take place when the values of σ and τ
lie below the curve AB. Because there are now three unequal principal stresses, the
figure shows that the value of the intermediate principal stress σ2 does not really
affect failure and the need to use σ3 becomes evident to draw the failure envelope.
Failure will take place if the circle with center C on σ1 and σ3 as a diameter just
touches AB. The Mohr envelope is usually concave downward, so that as the mean
stress ½(σ1 + σ3) is increased the plane of fracture becomes inclined at an increasing
angle to the σ1 direction. The point at which the Mohr envelope intersects the τ axis
is the shear strength of the rock. It should also be emphasized that in petroleum
reservoirs, the pore pressure needs to be considered because failure occurs when
the matrix stress reaches a critical level. Therefore, the effective stress is normally
used when possible. Fig. 16 is a typical Mohr failure envelope illustrating several
tests.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 27 of 52

Fig. 16. Typical failure envelope.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 28 of 52

4 Measurement of Rock Properties


Rock properties can be measured using both core analysis techniques and logging
techniques; Fig. 17 lists these properties for review.

ν Poisson's Ratio Lateral strain


Longitudinal strain

G Applied stress
Shear Modulus
Shear strain

E Applied uniaxial stress


Young's Modulus
Normal strain

Kb Hydrostatic pressure
Bulk Modulus
Volumetric strain

Cb Bulk Compressibility (with porosity) Volumetric deformation


Hydrostatic pressure

Cr Rock Compressibility (zero porosity) Change in matrix volume


Hydrostatic pressure

α Biot Elastic Constant Pore pressure proportionality

Fig. 17. Dynamic elastic properties.


Generally, it is said that core analysis would be used to obtain the “static” data while
logging would give the “dynamic” data for direct stress-related parameters such as
the Young's modulus and Poisson's ratio. The core laboratory can also obtain
dynamic data by use of an acoustic wave. The sets of data can now be compared
and used to calibrate the logging method to obtain equivalent static data. The
reason calibration is important is that a fracture in rock will propagate in a quasistatic
regime and not at the speed of sound. Therefore, it is very important to understand
these differences and to distinguish the static from the dynamic because the
magnitudes are very different. In general, the values of the elastic constants
obtained by static techniques are lower than those obtained by dynamic methods.
Table 2 is taken from a rock mechanics report and shows values from static and
dynamic tests. This difference is probably due to the presence of microcracks, which
remain unaffected by the frequencies associated with sonic methods.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 29 of 52

Table 2. Dynamic Versus Static Tests


Depth (ft) Dynamic Dynamic Static Static Compressive
Young's Poisson's Young's Poisson's Strength (psi)
Modulus Ratio Modulus Ratio
6 6
(X 10 psi) (X 10 psi)
4728H 4.7 .28 4.3 .27 32,200
4728V 4.7 .18 4.1 .16 29,500
4731H 5.3 .27
4731V 5.5 .26
4737H 4.0 .33 28,000
4737V 2.6 .17 30,300
4738H 7.7 .24
4738V 4.9 .19
5434H 4.0 .26
5434V 4.8 .12 5.3 .19 42,700
5435H 4.3 .26 3.5 .38 31,800
5436H 3.5 .29
5436V 3.4 .17
5443H 6.0 .22
5443V 3.7 .12
5445H 7.0 .18
5473H 5.8 .26 3.5 .10 42,100
5473V 6.2 .21 3.4 .15 31,500

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 30 of 52

Uniaxial and triaxial tests are the most useful and common tests to determine the
mechanical properties of rock. The difference between these two tests is simply the
presence or absence of a confining pressure applied to the rock. A typical triaxial
test uses a cylinder of rock where the confining pressure is applied around the core
(axisymmetric) and the load is applied in the longitudinal direction. The confining
and longitudinal loads should be the same as the in-situ state of stress in the
formation. A series of tests is performed using a different stress and pore pressure,
and information such as deformation versus load is used to obtain the Young's
modulus and Poisson's ratio as covered in Section 3.4. This review and the previous
discussion on pore pressure show the importance of the stress; therefore, the need
to duplicate in-situ stresses is very important in the laboratory testing if accurate data
are to be generated for the critical rock properties. Methods to obtain the in-situ
stress data for use in the rock mechanics laboratory will be covered later.

4.1 Typical Properties


Some typical rock properties that are of interest when designing hydraulic fracturing
treatments are the porosity, permeability and elastic constants such as the Young's
modulus and Poisson's ratio. Compressibility and fracture toughness also enter into
the problem of optimizing the fracture design. Each one of these will be discussed
briefly so that a comparison can be made between rock mechanics laboratory data
and logging or other methods.
The two rock properties that are not measured directly using stress components
alone are porosity and permeability. These two properties are discussed as follows.
• Porosity — Information on the porous nature of the rock is necessary to use as a
guide to mechanical performance. The presence of pores in rock will decrease
the strength and increase the deformability. Porosity is the fraction of the bulk
volume of a material that is not occupied by solids; therefore, it is a measure of
the storage capacity of the reservoir. Samples submitted to the laboratory for
porosity determination should be coherent rock that does not swell or disintegrate
when oven-dried or immersed in water.
Porosity can be calculated by either water saturation or gas saturation after a
known volume of rock has been thoroughly desiccated. Some error is
introduced, particularly in the water saturation, because of the lower limit in pore
size that the saturating fluid can penetrate. Water-sensitive clays present in the
pores can also lead to erroneous results. The degree of saturation is improved
with gas, but one must question the application of gas-determined porosity (or
permeability) if the reservoir fluid is oil or oil and water combinations. Remember
that porosity can be strongly affected by the application of confining stress, so to
be realistic the porosity needs to be determined under simulated downhole
conditions. A porosimeter capable of simulating in-situ conditions should be
used, or data determined from triaxial tests can be used to estimate the influence
of stress.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 31 of 52

Porosity obtained from logging methods is covered in the lithology portion of this
section.
• Permeability — The same relationships using Darcy's law discussed in the
section on Reservoir Evaluation also apply to calculating permeability from core
measurements. Again, permeability is very dependent on the in-situ state of
stress and should be determined under simulated field conditions. Totally
duplicating in-situ conditions becomes very difficult because of the inability to
obtain the saturation phases and, therefore, not allowing the calculation of
effective permeability. The effective permeabilities depend not only on the rock
itself but also on the relative amounts and properties of the different fluids in the
pores.
Secondary permeability in the form of small fractures can also have a
tremendous impact on laboratory-obtained permeability. Rock containing natural
fractures must be tested under in-situ conditions; the signature of these fractures
can be seen in the nonlinear response of the confining pressure versus
permeability. Fig. 18 shows this relationship and is really similar to the effects
shown in Fig. 10 (from O to A).
Logging methods to obtain permeability include many assumptions and details
that cannot be covered here. However, correlation charts and equations are
available relating the resistivity gradient to permeability as a function of the oil
gravity, water density and 100% water-saturated formation resistivity. Other
techniques use a relationship between porosity and the irreducible water
saturation, but these are more general empirical relationships and may not exhibit
general application or validity. A typical field technique is to establish a
permeability versus porosity relationship from core testing. This relationship can
then be applied to the log porosity to obtain the log permeability. The best
method to obtain the permeability is still the well performance test.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 32 of 52

Fig. 18. Permeability versus confining stress for a fractured rock specimen.
The values for Young's modulus and Poisson's ratio, as mentioned earlier, should be
the static values for use in the fracture geometry programs. Along with
compressibility and Biot's constant, these are known as the elastic constants.
Young's modulus is important to calculate the net fracturing pressure and width
profile. Differences in the Young's modulus between the pay zone and the bounding
layers may restrict fracture growth if the modulus of the bounding layer is greater
than the modulus of the pay zone. This effect is due to the width being smaller in the
high modulus material, causing flow resistance and making fracturing more difficult.
Poisson's ratio is used primarily for calculation of fracture-width distribution.
Dynamic elastic constants are derived from the measurement of the elastic wave
velocities in the rock. Sonic logging and waveform analysis provide the means to
obtain continuous measurement of the compressional and shear velocities. Using
these data, in conjunction with a bulk density measurement, allows the in-situ
measurement and calculation of the mechanical properties of the rock. Problems
arise when shear travel time cannot be measured (soft formations or poor cement).
However, new tools such as the Dipole Shear Sonic Imager (DSSI) allow more
accurate measurement of the shear travel time in soft formations. Also, the
relationships of the dynamic elastic properties do not account for the influence of
fluid type on the sonic log response. The reader should examine the literature and
become familiar with the significance of these differences.

4.2 Toughness
Fracture toughness has been a controversial subject for some time now as different
people argue not only the impact that it has on fracture geometry, but also the actual
value of the number and the way it should be calculated. These questions will not be
discussed here, and the reader can pursue the problem by learning more about the
Griffith crack theory and investigating literature dealing with toughness as presented
by the various authors.
Fracture mechanics assumes that there are always pre-existing defects in a rock.
These defects induce high stress concentrations in the vicinity of the defect and
become the nucleus for crack initiation and/or propagation. Fig. 19 illustrates a
Griffith crack in a linear elastic medium.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 33 of 52

Fig. 19. Stress concentration near the tip of a crack.


Linear elastic fracture mechanics states that a fracture will advance when the stress
intensity factor, KI (opening mode), reaches a critical value, KIc (designated as the
fracture toughness). The magnitude of the stresses in the vicinity of a Griffith crack
follows an r-½ relationship—

K I = yσ πr (20)

where y is the shape function, r is the crack half-length and σ is the stress field. The
unit for fracture toughness is the pressure times the square root of the length.
The existence of large stress values near the tip of the crack make it very difficult to
develop a test configuration because of the swarm of microcracks created ahead of
the crack tip. This area of microcracks has to be limited when testing so that it does
not reach the edge of the laboratory sample. This area must also be relatively small
compared to the size of the crack if linear elastic calculations are to be valid. A
modified ring from a slice of core is used to test for fracture toughness. The slice
contains a central hole and two diametrically opposed flat surfaces as illustrated in
Fig. 20. The test is run under a constant axial displacement rate, and both the load
and the axial displacement are continuously recorded.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 34 of 52

Fig. 20. Modified ring.

Fig. 21 shows the curve associated with this test.

Fig. 21. Load/displacement curve obtained during a Modified Ring Test.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 35 of 52

The portion from O to A shows the stresses are reaching the critical intensity for the
crack to initiate. A to B is an unstable fracture propagation in the region of the high
tensile stress gradients. Beyond B the crack growth becomes stable. These tests
can be conducted under a confining pressure to simulate the actual field conditions.
A test under confining pressure will show an increase in fracture toughness values
when compared to unconfined tests. Some general concluding statements
pertaining to fracture toughness are
• Fracture toughness is a measure of the ease with which a fracture can propagate
in the rock. The higher the fracture toughness the more energy is used up in
extending the fracture, and the final fracture length may be significantly shorter.
• Contrasts in fracture toughness between rock layers have a significant effect on
fracture geometry. Bounding layers with higher fracture toughness than the pay
zone will act to contain the hydraulic fracture between them. A pay zone with
higher fracture toughness than the bounding layers will allow fracture growth in
the bounding layers at the expense of growth in the pay zone.
• The presence of natural fractures and the increased probability of the fracture to
follow a path of reduced resistance will tend to reduce the effective value of the
fracture toughness.
• For cases with relatively small pressure gradients in the fracturing fluid (large
fracture resulting from small elastic moduli or low-viscosity fracturing fluid), the
predicted fracture geometry can depend significantly on the value of the fracture
toughness.
• Propagation of fractures over long distances may increase the size of the process
zone, resulting in a larger value for toughness. Hydrostatic pressure has also
been shown to increase the value of the toughness. These items may explain
why the pressures observed at the wellbore are often larger than the pressures
predicted by fracture analysis.
• The fracture toughness for many competent rocks is in the order of 1000
(psi times square root of inches). This value is used many times as the default
value in fracture geometry models.
• Fracture toughness must not be confused with the tensile strength, although the
two properties are related.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 36 of 52

5 Determining In-Situ Stress


Some of the previous discussion has emphasized the need for accurate in-situ stress
data for use in running laboratory tests, and for calibrating certain logging
techniques. The total stress picture is influenced by lithology, tectonics, method and
type of burial (sedimentation), and the effect of erosion or glaciation. Many
assumptions are made in the methods used to calculate stress and in many cases
significant error may be introduced by the assumptions. The drilling of a borehole
will distort the pre-existing stress field. However, these induced stresses will
diminish away from the wellbore. Consequently, these stresses near the wellbore
(stress concentration) affect the pressure needed to induce a fracture but not the
propagation of the fracture away from the wall of the wellbore. Also, as mentioned
previously, rocks are nonideal. Anisotropy is common in many rocks, even without a
discontinuous structure, because of the preferred orientations of the mineral grains
or the directional stress history.
Bedding or layering alone can make shales, thin sandstones and limestones, and
other sedimentary rock highly anisotropic. Even rock that may be free of a bedding
structure, such as thick sandstone or limestone, can have directional properties
because of unequal principal stresses they may have been subjected to during the
transformation from sediment to rock. Any fissured rock has properties that are
greatly influenced by the state of the stress across the fissures. The rock behaves
as one material when the fissures are closed and as a different material when the
fissures are opened or sheared. The behavior of a network of fissures is more
important to the rock properties than the mineral composition. Fissures, as well as
pores and vugs, can create a nonlinear load and deformation response, reduce
tensile strength, create stress dependency in the material properties, and produce
variability and scatter in test results.
The total state of anisotropy is very difficult to duplicate for exact laboratory testing.
This is the reason that the critical stresses and elastic properties should be
determined with in-situ methods. Some of these methods include specialized core
tests, microfracturing, pump-in tests and certain logging techniques. It should be
pointed out that this concern and need for accurate in-situ data are simply to allow
an accurate prediction of the fracture azimuth and geometry for true optimization.
The data on stress are even more appreciated when working with three-dimensional
fracturing design or analysis models.

5.1 Core Tests


Anelastic Strain Recovery (ASR) is a method that requires oriented core, and the
measurements are made at the drilling site. A series of rings containing precision
gauges is immediately placed over the core when it is brought to the surface. Fig. 22
is a simple drawing of how one ring might appear. (Holcomb, D.J. and McNamee,
M.J.: Sandia Report SAND 84-0651, 1984).

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 37 of 52

Fig. 22. Drawing of the LVDT* displacement gauge developed at the


Sandia National Laboratory1
*LVDT = Linear Variable Differential Transformer
1
Holcomb, D.J. and McNamee, M.J.: Sandia Report SAND 84-0651, 1984.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 38 of 52

As many as 36 gauges might be used to measure the response from a core piece.
This test measures the relaxation which the core undergoes after detachment from
the stressed rock mass. Fig. 23 shows the simple basics of a typical ASR curve.

Fig. 23. Typical ASR curve.


An instantaneous elastic part is indicated from A to B and occurs as soon as the core
is cut. The time-dependent anelastic component is what is actually measured and is
shown as the difference from B to whatever the value is at the end of the
measurements. The actual test measures the strain in various directions, and the
principal strain axis are then determined. Fig. 24 shows an actual ASR curve.
These principal directions are assumed to be the same as the principal axis for the
in-situ stresses.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 39 of 52

Fig. 24. Anelastic strain recovery measurements.


Results are interpreted based on the assumption that because of the stress
relaxation, microcracks develop in the core and that the dominant microcrack set will
be oriented perpendicular to the maximum horizontal stress. The absolute
magnitude of the stress is determined by the relationship to the known overburden
stress.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 40 of 52

Differential Strain Curve Analysis (DSCA) is another approach that relies on strain
relaxation, and at first glance seems to be based on the same fundamentals as ASR.
Oriented core is also needed for DSCA, and the test relies on the assumption that
the distribution and density of the resulting microfracturing that occurs are directly
proportional to the stress reduction that the core has sustained. This test uses a
cube machined from a core, and an array of strain gages is cemented to the sample
as shown in Fig. 25.

Fig. 25. Gage pattern.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 41 of 52

The cube is encased in a flexible silicone jacket to prevent fluid migration into the
pores, and the sample is then hydrostatically compressed (stress is applied equally
in all directions) at a constant rate. During loading, and the subsequent unloading,
strains and pressure are recorded every second. Several cycles are typically
performed to evaluate the effects of nonelastic crack closure. At the beginning of the
loading cycle, the existence of microdiscontinuities (such as microcracks) introduce a
softening. The very accurate strain measurements in the various directions make
possible the characterization of the microcracking, which is related to the
measurement of the pre-existing state of stress. Fig. 26 shows a typical DSCA
curve.

Fig. 26. Typical DSCA plots.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 42 of 52

The compressibility of the rock as a function of the pressure can then be calculated
by derivation of the strain/pressure curves. The principal strains and the respective
directions are then computed, and from this the stress ratios and stress directions
are determined and presented graphically on a stereoplot. Fig. 27 is an actual
stereoplot from a DSCA test on a core from 8114 ft. The figure shows the azimuth of
least stress to be NE/SW; therefore, an induced hydraulic fracture should propagate
approximately at right angles to that direction.

Fig. 27. Stereoplot.


The main differences between ASR and DSCA are that ASR data are limited to the
present state of the stress, whereas DSCA is a reflection of the total stress history of
the rock. Both cases still require certain assumptions for complete analysis.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 43 of 52

5.2 Microfracturing
This is a test where a very small volume (100 to 500 gal) of low-viscosity fluid is
injected into a well that is perforated at the point of test interest. The objective of the
test is to determine the in-situ minimum horizontal stress. Flow rates are usually in
the 10- to 15-gal/min range, but this will depend on several other conditions (such as
perforations and interval size). After a small volume of fluid has been injected, the
system is shut in and an accurate reading of the instantaneous shut-in pressure
(ISIP) is taken. The pressure is allowed to bleed off and another test, or cycle, is
started using the same fluid and pump rate as the first. The difference between the
breakdown pressure in Cycle 1 and the reopening pressure of Cycle 2 is an
indication of the tensile strength. The test can be repeated for several cycles,
looking for uniformity in the later data.
Fig. 28 illustrates some of the characteristics of an idealized pressure plot from a
microfrac. The entire purpose of the test is to obtain an accurate value for ISIP
because it is generally accepted that σHmin= ISIP.

Fig. 28. Idealized plot.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 44 of 52

A microfrac test assumes one of the principal stresses is parallel to the axis of the
borehole (overburden stress). Obviously, this stress cannot be the minimum stress
or the test would simply determine the overburden pressure. Therefore, the test
must be made below the critical depth where horizontal stresses are not greater than
the overburden. The test interval should be well isolated with straddle packers or
packer and bridge plug, and a very high-quality pressure gauge (quartz crystal
oscillator) should be used to obtain accurate bottomhole pressure. Downhole
closure tools are available that allow a wireline pressure gauge assembly to seat on
a sub. This method provides downhole shut-in, yet positions the gauge close to the
test interval immediately after pumping. Fig. 29 is an example of the pressure and
injection rate data from an actual microfrac. Most microfrac tests show that the
fracture reopening pressures tend to decline with successive pump cycles.

Fig. 29. Microfrac test.


Sources of error for these tests include the effects of leakoff (why small volumes are
used), pore pressure, stress concentrations and rock strength. If pore pressure is
altered during the test period, the shut-in pressure could possibly increase. High-
viscosity fluids may become necessary in some cases where leakoff with a thin fluid
is excessive. However, in the application for zone breakdowns, it should be
remembered from the earlier discussion of the effective stress law that an increase in
pore pressure in the vicinity of the wellbore will decrease the breakdown pressure.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 45 of 52

5.3 Pump-In/Flowback Test


This test consists of a pump-in portion and a controlled flowback portion and is used
to determine closure pressure for application in the DataFRAC* Service. A step-rate
test is sometimes incorporated as a part of the pump-in test. The step-rate test is a
pump-in where an initial low injection rate is increased at a constant time step. A
Cartesian plot of rate versus pressure should show a break in the curve where the
fracturing of the rock begins. Continued pumping will not increase the pressure at
the same magnitude as prior to the fracturing. The intersection of two straight lines
drawn through the two portions of the curve is used to determine the extension
pressure, but not the closure pressure. Once the maximum rate above the fracture
pressure is reached and maintained for a short time, the injection is stopped and a
controlled flowback is performed to obtain the closure pressure. Several pump-
in/flowback tests can then be performed by starting injection at the previous
maximum rate used in the combined test followed again with a controlled flowback.
These tests are discussed in more detail in DataFRAC Service, and are briefly
discussed here only to point out the differences and the fact that the stress is not the
same as that calculated from a microfrac test.
The minimum stress, σmin, is considered to be a local stress and generally varies
through the gross pay zone and even through any bounding layers that may or may
not be barriers. Closure pressure is derived after using larger fluid volumes, to
create a fracture in the entire zone, and is a global parameter defining the fluid
pressure for the fracture closure without the proppant. Therefore, the closure
pressure is an average of the formation characteristics and provides the bulk
property of the formation to be fractured. Also, because small fluid volumes are
used in a microfrac test, the net pressure at shut-in is very low and the shut-in
pressure can be used to infer the stress. The net pressure generated during the
procedure to determine closure pressure can be relatively large at shut-in.
Therefore, the pressure at the wellbore is not the same as the closure pressure.
This closure pressure must be determined by other means that are well documented
in the references.
It must also be pointed out that the key to a successful DataFRAC interpretation is
by independently obtaining the closure pressure. Parametric analysis using the
DataFRAC program can then be used to compare the values of the Young's
modulus, fracture height and several other parameters. This is why it is so important
to have good rock mechanics data for comparison.

5.4 Logs
The previous discussions on the dynamic properties obtained by logging have
already set the stage for this discussion. The difference between dynamic data and
static data is still a major concern when trying to directly apply log-derived data to
fracture design and analysis problems. Much improvement is being made by

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 46 of 52

computer programs that present the stress profile as accurately as possible by using
built-in correlations obtained from field tests. Presently, logs are the only way to
obtain a complete stress profile. It must be emphasized that these logs need to be
calibrated with values obtained from other methods.
Obtaining and processing sonic data is the basis for logging systems that provide
stress information. Basically, two waveforms are analyzed to provide the data.
These waveforms are the P waves (compressional) and the S waves (shear).
Computer analysis processes the data and produces the log presentation. Early
mechanical properties logs presented three tracks that gave the Young's modulus,
Poisson's ratio and fracture gradient. Standard sonic tools were used and the
greatest problem was the inability to process the shear wave for analysis. Improved
techniques and new tools, such as the dipole shear sonic imaging, and the long-
spaced sonic and sonic digital tools, now make it possible to obtain the independent
determination of the P and S waves during wellbore logging. This in turn allows the
calculation of stress and mechanical properties every six inches over the logging run.
The fracture-height log is a presentation of the stress profile and can be used to
model fracture-height growth.
The current procedure used to generate the fracture-height log includes
• compressional and shear waves are used to calculate the elastic properties of the
rock
• an elastic model is used to compute the minimum horizontal stress
• a linear fracture mechanics model is used to predict the fracture-height growth.
The log that is produced from this analysis is shown in Fig. 30.
The first track on the left gives the depth and location of the perforated interval(s).
The second track is the gamma ray. The third track (delta pressure) is given in
200-psi increments, and is the net pressure required for the fracture height to grow
vertically. When more than one zone is being fractured at the same time, the
program calculates the percentage of fracturing fluid that enters each fracture based
on the zone thickness and material balance. The remaining tracks to the right are
the Young's modulus with the minimum horizontal stress gradient, and Poisson's
ratio.
Several things are obvious from a quick look at this log. The net pressure track is
divided in 200-psi increments, so looking at the bottom set of perforations shows
200 psi more is needed to initiate a fracture here compared to the upper zone. If the
net pressure does not increase by 200 psi and allow the pad to develop a fracture in
the bottom zone, then the proppant could cause an early screenout in the bottom
zone and the entire treatment would then be placed in the upper zone. At the same
time, note the fracture-height growth in the upper zone before the lower zone
fractures. At 600 psi the fracture height is excessive. Therefore, the ability to control
or limit the net pressure becomes an important consideration to obtain an efficient
fracturing treatment. Having this information allows the engineer to model and
compare various injection rates, fracturing fluid viscosity and other parameters that

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 47 of 52

control the net pressure. This situation is just one of many that may justify doing a
ballout job prior to the fracturing treatment. Again, this log should be calibrated by
pump-in tests whenever possible.

Fig. 30. Fracture-height log.


Fracture-height prediction methods were not very accurate in the past, and at best
only qualitative until the late 1970s. A very common method was to use the gamma
ray (GR) and the spontaneous potential (SP) to differentiate sands from shale. The
location of the shale was considered to be a limit for the fracture height in that
direction. The misconception that shale is an absolute barrier to fracture-height
growth still exists today. Looking again at Fig. 30 shows that there is really no way
to determine the stress by simply examining the gamma ray.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 48 of 52

6 Fracture Height (Post)


Determination of the fracture height after fracturing is briefly discussed in this section
because of the importance for calibrating tests and logs. Fracture height has always
been very difficult to obtain and yet is so important for both design and evaluation.

6.1 Radioactive Tracer


The use of a single tracing element brings up the question of whether to run it with
the fluid or use it in the proppant only. Tracer used in the fluid will indicate the total
fracture height, while tagging of the proppant will show the proppant height. The
ideal method is to use two different isotopes and tag both the fluid and the proppant
so the height profile will show the location of both the fluid and proppant. In some
cases, the difference between the fluid height and proppant height is an indication of
the width of the fracture. Obviously, bridging had to occur or the proppant height
would equal the fluid height. This not uncommon finding allows a better picture of
the elastic properties of the rock that was fractured, and may also help the calibration
of the logs. Carrying this concept one step further is to use three isotopes, two for
fluid and initial proppant, and the third for any tail-in proppant. Many times it is
apparent from this tracer combination that the tail-in proppant has significantly less
height than the initial proppant. This situation may explain why the transmissibility
obtained from well performance tests does not match what the design used when
assuming the tail-in proppant would cover the entire net pay.
Each isotope emits a unique combination of gamma rays. The newer, natural,
gamma ray spectroscopy tool can measure the count rate versus depth for multiple
isotopes used in a fracturing treatment. Fig. 31 is a two-isotope tracer log where one
tracer was used in the lower zone treatment, and another was used in the upper
zone treatment (two-stage fracturing treatment).

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 49 of 52

Fig. 31. Two-isotope tracking of a two-stage fracture treatment.


The purpose of the multiple isotopes in this case was to determine if the upper zone
fracture would propagate into the lower zone. The lower zone was treated using
scandium. The trace is shown in black, as well as by the difference between the
before and after gamma ray. Note that the gamma ray traces come together above
the zone but continue with some separation below the zone, indicating more
downward than upward growth. Also, the black extending up into the top zone may
be contamination on the pipe. This pipe contamination can be minimized by cutting
the tracer material 10 to 50 BBL before all of the propping agent is pumped so the
remaining propping agent will scour the pipe. The iridium trace on the upper zone
indicates the upper fracture did not communicate downward to the lower zone.
However, there is not enough log above the zone to see the height growth above.
Remember that the depth of the logging tool investigation also has a big influence on
the presentation and interpretation of a tracer survey.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 50 of 52

The next example uses two isotopes in a single-stage treatment, and the tracer log is
shown in Fig. 32.

Fig. 32. Two-isotope tracking of a single-stage fracture treatment.


The pad and initial proppant were tagged with scandium, while the remainder of the
proppant was tagged with iridium. The log shows no indication of scandium
(assuming it was pumped), because it must be out of the radius of the tool
measurement. Several interpretations could be made from this unusual situation.
First of all, a tracer usually leaves some signature whenever it enters the formation.
The total absence of scandium might indicate that the fracture-height growth was
small, at least during the portion of the treatment containing the scandium, and then
grew to the indicated position later in the treatment. Examination of the net pressure
plot would help interpret this situation. Secondly, because the fluid was not tagged
in the latter part of the treatment, the iridium is an indication of the proppant only.
Total fracture height from this presentation can only be assumed from after the
gamma ray, where contaminated fluid (radioactive) is indicated above and a little
below.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Rock Mechanics
Page 51 of 52

6.2 Temperature
Temperature logs used for the purpose of determining fracture height is probably an
art as much as it is a science. There is not space enough here to cover the subtle
points of temperature logging. However, certain concepts that need to be
considered are
• A temperature log is measuring nothing more than the temperature of the fluid
filling the well. The thermal conductivity of the various rocks will affect a
temperature profile after pumping. Therefore, an accurate prefracture profile can
be run by circulating fluid as close to the fracturing rate as possible and then
running the base log.
• Low bottomhole temperature wells may not have enough contrast with the
injected fluid temperature to show the sufficient differences after fracturing.
• Temperature surveys will not indicate the proppant height but rather the total
fracture height. Frequent temperature runs comparing the changes in
temperature can be used to obtain a qualitative fracture width.
• Proppant fill below the perforated interval prevents the tool from reading below
this depth. The temperature anomaly will be lost by the time the wellbore is
cleaned out.
• Any fluid movement prior to temperature logging or during logging may cause the
interpretation to be difficult or even impossible. Care should be taken to prevent
fluid from flowing back prior to or during logging.
• Because the tool is measuring the temperature of the fluid in the pipe, there may
be a portion of fluid below the perforations (and below the turbulence level
caused by pumping) where the original bottomhole temperature is still
represented. The tool entering this region below the perforations will give the
typical kick to the right showing heat-up to the base curve. Many times this is
erroneously picked as the bottom of the fracture. A fracture that is in this area
must have some time to influence the temperature inside the pipe. This is only
seen when successive temperature runs are made and a crossover is finally seen
below the perforations, indicating that the fracture finally had a temperature
impact on the casing fluid.

DOWELL CONFIDENTIAL
Section 200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Rock Mechanics Dowell
Page 52 of 52

Fig. 33 is an example temperature survey with a backup radioactive tracer.

Fig. 33. Temperature survey with RA tracer.


It is not known if the radioactive material was used to tag the fluid or proppant, but
comparing the upper gamma ray portion with the temperature log may indicate that
the isotope was in the proppant. It is also interesting to note the rapid increase in
temperature below the perforations, yet the tracer still shows gamma ray separation
below the bottom of the fracture as marked on the temperature profile.
It must be understood that both temperature and radioactive logs used to determine
the fracture height are a “best case” presentation. The assumption is that the
vertical fracture is totally present within the wellbore. When either the wellbore is
inclined to the fracture plane, or the fracture plane is inclined to a vertical wellbore,
then the tool will only read the portion in the wellbore. For example, this is important
when comparing the postfracture height and the height calculated using the
DataFRAC software.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 1 of 35

FRACTURE MODELING

1 Introductory Summary............................................................................................................. 2

2 Concepts................................................................................................................................... 3
2.1 Fundamental Laws............................................................................................................... 4
2.2 Constitutive Laws ................................................................................................................. 4
2.3 Fracture Propagation ........................................................................................................... 6

3 Hydraulic Fracturing Models .................................................................................................. 9


3.1 Two-Dimensional (2D) ....................................................................................................... 11
3.2 Pseudo Three-Dimensional (P-3D) .................................................................................... 15
3.3 Planar Three-Dimensional (PL-3D).................................................................................... 18
3.4 Fully Three-Dimensional (3D) ............................................................................................ 19

4 Examples ................................................................................................................................ 20
4.1 Case History ...................................................................................................................... 20
4.2 Model Comparisons ........................................................................................................... 28
FIGURES
Fig. 1. Modes of loading............................................................................................................... 7
Fig. 2. Fracture divided into elements.......................................................................................... 9
Fig. 3. Representation of a planar fracture. ............................................................................... 10
Fig. 4. KGD geometry. ............................................................................................................... 11
Fig. 5. PKN geometry................................................................................................................. 12
Fig. 6. 2D and radial Sneddon cracks. ....................................................................................... 13
Fig. 7. Elliptical profile (P-3D)..................................................................................................... 17
Fig. 8. Example grid (PL-3D model)........................................................................................... 18
Fig. 9. Fracture profile (PL-3D model). ...................................................................................... 19
Fig. 10. Permeability, thickness and stress profile. .................................................................... 20
Fig. 11. Computed values for Young's modulus and Poisson's ratio. ........................................ 21
Fig. 12. Profile of bottomhole, casing and tubing pressures. ..................................................... 24
Fig. 13. Pressure match for bottomhole and casing pressure. .................................................. 24
Fig. 14. Fracture profile.............................................................................................................. 25
Fig. 15. Fracture width profile. ................................................................................................... 25
Fig. 16. Match of net pressure for calibration fracture and main fracture. ................................. 26
Fig. 17. Fracture profile.............................................................................................................. 26
Fig. 18. Reservoir model for final history match......................................................................... 28

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 2 of 35

Fig. 19. TRIFRAC length and width profile. ................................................................................32


Fig. 20. STIMPLAN length and width. ........................................................................................32
Fig. 21. FRACPRO length and width profile. ..............................................................................33
Fig. 22. GOHFER length and width profile. ................................................................................33
Fig. 23. TERRAFRAC length profile. ..........................................................................................34
Fig. 24. STIMPLAN length and width profile...............................................................................34
Fig. 25. MEYER length and width profile. ...................................................................................35
Fig. 26. Ohio state length profile.................................................................................................35

TABLES

Table 1. Comparison Of Stress ..................................................................................................22


Table 2. Permeability and Fluid Loss .........................................................................................22
Table 3. Design Information .......................................................................................................23
Table 4. Fracture Model Comparison Runs................................................................................30
Table 5. Fracture Model Comparison Runs................................................................................31

1 Introductory Summary
The prediction of fracture geometry has been a central issue in engineering design
and evaluation of hydraulic fractures, and many models have been developed over
the years. These models determine fracture geometry by attempting to relate many
variables such as rock properties, fluid properties, fluid volume pumped and stress
data. Some models use a fixed fracture height and others continuously calculate the
height during the simulation. Each change aimed at more closely matching the real
conditions requires more sophistication in modeling the fluid flow in the entire
fracture, effect of proppant and elasticity of the entire system. To be practical,
however, the calculations must be made at reasonable increments along the fracture
and computational time must not be excessive. The degree of sophistication of a
model is therefore somewhat controlled by the practical application. The models are
also data limited.
The comparison of different models can be difficult and confusing because of the
way the various authors handle the variety of conditions, what they feel is important,
what assumptions they make and how portions of the model are coupled. Decisions
on how to handle elasticity, fluid flow, type of grid or cross section, vertical stress
differences and toughness, for example, can have a large impact on the calculated
fracture geometry. There is still much work to be done in obtaining meaningful data
for input into the more sophisticated models.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 3 of 35

Basically, there are four types of fracture models either being used or being
developed in the industry today.
• Two-Dimensional (2D)
• Pseudo Three-Dimensional (P-3D)
• Planar Three-Dimensional (PL-3D)
• Fully Three-Dimensional (3D).
The designation of the third listed model as “Planar” is an area of confusion that
should be clarified. Actually, the term planar means that the fracture occurs in a
plane. This condition is true for all fracture models except the fully 3D. Planar was
simply used to name the model that is more advanced than the pseudo 3D, but not
quite as sophisticated as the fully 3D. A fully 3D model would have the capability of
being nonplanar (fracture could curve or change planes) if the correct stress data
and other information were available for input.
This section on fracture models will be limited to a brief discussion of some of the
concepts that must be considered to build a model, as well as a brief discussion of
each model. It is beyond the scope of this section to cover each model in detail
because of the number of models available, and not having the documentation or
code to examine each model. Also, models change as more data become available
from evaluation, in-situ testing and calibration of logs, and from special industry
projects to calibrate the various models based on the best available information. A
successful model is one that has the ability to match the pressure from the treatment
by using realistic variables based on in-situ data, and to calculate a fracture height
consistent with other methods used on the actual treatment.

2 Concepts
Modeling fracturing treatments requires a blending of many different components,
such as rock mechanics, fluid mechanics, rheology and heat transfer. Two sets of
laws are required for this process.
• The Fundamental Laws dealing with mass, momentum and energy conservation.
These relate to the physical principles.
• The Constitutive Laws include rock elasticity and fluid rheology. These describe
the behavior of a system under a certain number of conditions.
Coupling these two sets with the appropriate “boundary conditions” produces some
very complicated mathematical formulations. To solve the coupled problem requires
discretization of the system (break into small geometric components such as a grid),
and then writing equations in a form that can be solved with digital computing.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 4 of 35

2.1 Fundamental Laws


• Conservation of Mass: The mass of a system does not change with time. The
conservation of mass for modeling a fracture treatment is used to give the overall
material balance. This material balance is illustrated by Eq. 1.
Vi = V fp + VLP (1)

Where:
Vi = volume injected
Vfp = fracture volume
VLp = injected volume lost to the formation.
• Conservation of Momentum: Two types of forces can be distinguished; (1) the
body forces such as gravity that act on the whole volume, and (2) the surface
forces, such as pressure forces and fluid friction, that act only along the boundary
of the domain.
The conservation of momentum principle relates the time change in the total
momentum of a body to the applied forces (both on the volume and on the
surface). Many fracture treatments are modeled as quasistatic. This implies that
the rates of the change of velocities are negligible, and therefore the summation
of surface and volume forces is zero.
• Conservation of Energy: This pertains primarily to the two fundamental laws of
thermodynamics for a system; (1) the change in total energy of a system is equal
to the work of the forces applied on the system plus the rate of heat transfer, and
(2) the internal energy of a system is a function only of its entropy.

2.2 Constitutive Laws


A mechanical system is completely defined by a certain number of variables that
depend on time and position.
• mass
• temperature
• velocity (three components)
• stresses (six components).
These variables make a minimum of 11 unknowns while the number of components
in the three conservation laws is only five. Therefore, constitutive equations to solve
the system need to be defined. Some equations used in the less complicated
models can be simplified to the point where they no longer apply, thereby reducing
the number of computations needed to complete a simulation.
Three types of constitutive relations are considered for the system.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 5 of 35

1. Fluid incompressibility is assumed.


2. A relation exists between stress and strain (or rates of strain).
3. Scalar (magnitude only) quantities are related to fluxes.
• Incompressibility: For many fluids, the density does not change with time,
pressure or temperature. This assumption of constant density (incompressible
fluid) causes the equation of mass conservation to simplify and become zero.
• Stress/Strain: The section on rock mechanics showed the relationship between
stress (σ) and strain (∈) for an elastic solid, giving the coefficient of proportionality
(σ = E∈) where E is Young's modulus. Modulus contrasts act mainly to alter the
shape of the fracture rather than the tip positions at any pressure. Analyse to
determine the effect of modulus contrast are extremely time consuming because
they require large finite-element or similar solutions. Formulas for the average
width are used in many cases to scale the width profile.
For the general case of a fracture growing into the surrounding layers of greater
stress and/or fracture toughness than the pay zone, the pressure, height and
shape of the fracture cross section depend on the stress, toughness, thickness
and modulus of the individual layers. The more simple models would assume a
constant modulus in all layers, plane strain elasticity and no flow-induced
pressure drop in the vertical direction.
When considering fluid flow, the stress tensor includes pressure and a viscous
tensor (τ). Newton's law for viscous incompressible fluids is simplified to
τ = µγ (2)

where µ is the viscosity of the fluid and γ is the strain rate tensor. Flow in several
of the fracture models is assumed to be in one direction only (x direction) and the
velocity field is therefore unidirectional. This means that both the shear stress
(τxy) and rate of deformation: γxy are related as:

τ xy = µγxy (3)

• Coupling conservation relations with the constitutive equation leads to a very


complicated fluid equation. The more sophisticated PL-3D models may attempt
to handle this equation while the 2D and many P-3D models must make
simplifying assumptions, particularly on velocity, before numerically solving the
system. Handling fluid flow in the fracture to determine the pressure distribution
is very important and necessary to determine the fracture displacement, as well
as using a method to determine how a particular model will handle all of the
variables associated with adding the proppant to the system. An accurate
calculation of the pressure distribution within the fracture is critical for any realistic
simulation.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 6 of 35

• Flux Laws: These include the relationships of pressure drop in a porous medium
to velocity (Darcy's law), of heat flux to temperature (Fourier's law) or the rate of
reaction to the change in concentration (Fick's law) for acid fracturing.

2.3 Fracture Propagation


Modeling the fracturing process is a very complicated task. Injecting fluid into the
formation will modify the stress distribution and pressure within the formation,
creating a fracture in which the injected fluid flows. This injected fluid exchanges
heat with the formation and leaks off into the formation. The poroelastic effect can
be interpreted as a time-dependent back-stress; whenever a fracture profile is
obtained and a pressure is calculated, an additional term must be added to the
pressure. The fluid is also considered to be a multiphase type because of the
proppant being carried. The proppant also alters the fluid viscosity, which affects the
fluid-flow model. A good model needs rigorous calculation of the fluid fronts as well
as handling the settling of the proppant as in the simpler 2D models up to convection
hindered settling and other more complicated accounting of the proppant in some of
the PL-3D models.
Simulating the propagation of a hydraulic fracture requires consideration of the linear
elastic fracture mechanics (LEFM) formulation, fracture fluid flow and continuity
equation. These three sets need to be coupled to simulate the propagation; the
mathematical problem is complex because of the different types of equations and the
presence of a moving boundary (fracture edge).
The strains produced in the formation caused by the deformation from inducing a
hydraulic fracture are actually quite small. The small value of strain allows the
assumption that the formation deforms in a linear elastic manner. Knowing the state
of stress (or pressure) induced by the fluid in the fracture and the confining stresses
(boundary conditions) allows the calculation of the fracture width (displacement).
One simple method is to consider the crack as a uniformly pressurized elliptical
surface with semiaxes a and b. This condition forces the fracture to assume an
elliptical shape (Sneddon). Different cases may be derived from Sneddon's solution.
These cases can relate to the Griffith crack as well as the radial or penny-shaped
crack. This will be covered in more detail in the discussion of 2D models. The
classical techniques in elasticity do have severe limitations because of the stress
singularity near the crack tip, and special methods have to be used to improve the
accuracy in computing the stress intensity factors. Boundary integral techniques
have become popular because the problem is solved only at the boundary, reducing
the dimension of the problem by one, and provides a simple method to determine
stress intensity factors. However, implementation can be difficult for
nonhomogeneous media with varying elastic moduli.
Remember from the section on rock mechanics that the stress intensity factor (KI) is
for the opening mode of the fracture and describes the magnitude of the stresses
near the crack tip — it depends on the crack geometry, surrounding material and
applied loads. Fig. 1 shows the three fundamental modes of loading for a fracture.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 7 of 35

Fig. 1. Modes of loading.


The opening mode corresponds to normal tension, in front of the crack. The sliding
mode is associated with transverse shear while the tearing mode is a longitudinal
shear. Usually, only the opening mode applies for a plane fracture. The other
modes may be applicable in complex fracturing, out of plane fractures or when near
natural fissures.
For a given system of stresses, it is necessary to determine whether a fracture will
initiate and then propagate, and in which direction, and from which point in the
created fracture. Griffith (1921) addressed the problem by defining the surface
energy of a fracture. The Griffith criterion stated that there is an equilibrium if the
specific surface energy γ is related to the change in internal energy (U) by:
∂U
2γ = − (4)
∂A

where A is the fracture area. For a uniformly loaded fracture, the critical load (σc)
and the surface energy are related, such as
1/ 2
 2 Eγ 
σc =   (5)
 πx ( 1 − ν 2 ) 
 f 

where xf is the fracture half-length and E and ν are the elastic constants.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 8 of 35

Another important concept related to the surface energy criterion is that of the critical
stress-intensity factor (KIc), also known as fracture toughness. The condition for a
fracture to be in equilibrium requires that the stress intensity factor (KI) associated
with the load be equal to a critical value (KIc). A simple relation can be derived
between the surface energy and KIc for a uniformly pressurized fracture —
(1 − ν 2 )K 2 Ic
γ= (6)
E

This expression indicates (for linear elastic behavior) that the surface energy
criterion and the critical intensity factor are related and form a unified criterion for
propagation.
The calculation of the pressure distribution in the fracture, due to fluid flow, is
necessary to determine the fracture displacements. As mentioned earlier, the fluid-
flow problems to include non-Newtonian as well as Newtonian fluids present some
problems and require a large amount of computational time. Because of this
problem, the fracture may be discretized into a series of parallel lines (surfaces) and
the flow considered as quasistatic. The PL-3D models do handle the flow problem a
little more rigorously and, consequently, require more time for computation.
The continuity equation is the last relation and simply describes the conservation of
mass previously discussed. The continuity equation may be written for each fracture
element (in the grid) — Flow Rate In = Flow Rate Out + Accumulation.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 9 of 35

3 Hydraulic Fracturing Models


When a fracture is created in a rock mass, the fluid applies pressure to the two faces
of the resulting crack. This pressure causes the crack to open. The width at any
point is therefore dependent on the pressure applied to the faces. The condition of
the plane strain and rectangular elements is used to model the fracture for the 2D
cases. Fig. 2 illustrates a boundary element technique applied to a planar fracture.

Fig. 2. Fracture divided into elements.


The simplest representation is when the width is assumed to be piecewise constant
over each element. For problems solved by this method, the width distribution is
typically obtained for a specified distribution of pressure over the elements.
Alternatively, the pressure distribution can easily be found from a specified width
profile. Some P-3D models may use a little more sophisticated grid system, but fluid
flow is again only 1D and the vertical height growth is handled by shape factors and
other methods although elasticity is considered to be 2D. The PL-3D models use a
more complicated grid system (small squares or triangles). The junction of the grid
elements becomes the node for the solution of the problem. This allows the applied
pressure and displacements everywhere on the fracture faces to be related.
Elasticity is also 3D in these PL-3D models.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 10 of 35

The planar fracture was briefly introduced in the introductory summary to clarify a
model description. Fractures are planar features as shown in Fig. 3.

Fig. 3. Representation of a planar fracture.


The fracture lies in the xz plane, and the width is represented by the y direction.
Fluid may flow in all directions in this fracture, but 3D fluid flow is very complicated
and various models handle this flow differently. Usually flow along the fracture
length is modeled in detail, while horizontal flow across the width of the fracture and
vertical flow are neglected (2D, P-3D). Horizontal flow may be calculated in an
average sense by assuming that an average velocity adequately represents the flow.
An overall fluid-flow model could therefore be obtained by coupling 1D mass
conservation and momentum conservation equations. Volume conservation would
replace mass conservation if only incompressible fluids are considered.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 11 of 35

3.1 Two-Dimensional (2D)


Most 2D models require that a value for fracture height be input so the length and
width can be calculated from the volume and flow characteristics governed by the
code for the particular model. The models that will be discussed here are the (1)
Khristianovic and Zheltov, with later contributions by Geertsma and de Klerk (KGD
model), (2) Perkins and Kern, and later Nordgren (PKN model), and (3) the radial or
penny-shaped model.
A common simplifying assumption is that the lateral effects of a fracture are small
compared to the vertical effects and can be neglected. This condition is termed
plane strain and implies that each cross section acts independently of any other
section, so that the mechanical analysis need only be performed in two dimensions.
This plane-strain assumption is an integral part of the 2D fracture models. Two
cases of plane strain can be distinguished for the 2D formulations.
• Horizontal Plane-Strain Geometry: The fracture zone will deform independently
of the upper and lower layers. This will be possible for free slippage on these
layers and represent a fracture with horizontal penetration that is much smaller
than the vertical, and where the fracture shape does not depend on the vertical
position. This describes the KGD fracture geometry as shown in Fig. 4. This
geometry has a constant height with a cross section that is rectangular in shape.

Fig. 4. KGD geometry.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 12 of 35

• Vertical Plane-Strain Geometry: This condition exists for a large confinement


where the fracture is limited to a given zone. Each vertical cross section deforms
independently of the others. The fracture widths in the vertical direction are
coupled through the continuity and fluid-flow equations; however, because there
is no vertical extension (in each of the vertical sections) during simulation, the
pressure is uniform and the cross-sectional shape of the fracture height is
elliptical. This describes the PKN geometry as illustrated in Fig. 5.

Fig. 5. PKN geometry.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 13 of 35

Sneddon's solution for modeling the behavior of a fracture (for linear elastic
assumption) was mentioned earlier during the discussion of concepts. Fig. 6 shows
the 2D and radial Sneddon cracks. These solutions are for a 2D crack having one
dimension of infinite extent, and the other dimension of finite extent (d in Fig. 6). The
radial or penny-shaped crack is defined by the radius (R). The resulting width is
elliptical in shape for both types of cracks, and is proportional to one of the
characteristic dimensions (either d or R).

Fig. 6. 2D and radial Sneddon cracks.


The width is also proportional to the net pressure (pf - σmin) and inversely proportional
to the plane-strain modulus given as
E ′ = E / (1 − ν 2 ) (7)

Young's modulus and Poisson's ratio are at in-situ conditions (E is defined by the
tangent Young's modulus Et).
Sneddon's method has been used in different ways to model 2D fractures. The
characteristic dimension, d, is assumed to be the total tip-to-tip fracture length (2xf)
for the KGD model. Since d is assumed to be the total fracture length, then the
infinite dimension has to correspond to the fracture height. The other assumption is
that the characteristic dimension, d, is the fracture height. This is the condition for
the PKN model; since d is the fracture height, then the infinite dimension has to be
the fracture length.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 14 of 35

The KGD and PKN models consider the propagation of a vertical fracture of given
height. However, in some cases, the vertical stress is lower than the horizontal
stress and the fracture will propagate in a horizontal or inclined direction. Both
Perkins and Kern as well as Geertsma and de Klerk gave simplified expressions for
the propagation of a radial horizontal fracture. Similarly, vertical fractures may
propagate radially in thick formations where there are no barriers to height growth.
This situation leads to the same equations when the injection interval is small
compared to the fracture extension. The KGD model is valid when h>>xf. The PKN
model is valid when xf >>h. The radial model is most appropriate when the total
length (2xf or 2R) is approximately equal to the height. Again, the three sets of
equations to be coupled are the elasticity, continuity and fluid flow.
Some of the important characteristics and differences pertaining to the 2D models
are
• KGD Fracture Model

− A fixed fracture height is assumed, and fluid flow is horizontal only (in the
direction of the propagation).
− Crack opening is solved in the horizontal plane. As a result, the fracture
width does not vary with the fracture height, except by the boundary
condition set at the wellbore that specifies a constant total injection rate.
− Width is constant in the vertical direction because of the plane-strain
condition and individual horizontal planes.
− The model gives wider fracture widths and shorter fracture lengths when
compared to the PKN model.
− The flow resistance in the narrow rectangular vertical width is what
determines the fluid pressure gradient in the propagating direction.
− The excess pressure (net pressure) decreases with time, and in log-log
coordinates has a slope equal to -1/3.
− The model is most appropriate when the fracture length is smaller than the
fracture height.
• PKN Fracture Model

− A fixed fracture height is assumed and fluid flow is horizontal only (in the
direction of the propagation).
− Crack opening is solved in the vertical plane. As a result, the fracturing
fluid pressure is constant in vertical cross sections perpendicular to the
direction of the fracture propagation.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 15 of 35

− Each vertical cross section deforms individually and obtains an elliptic


shape with the maximum width in the center. The only coupling between
the different vertical cross sections is due to fluid flow in the fracture. There
is no elastic coupling between the planes.
− This model is used to describe the behavior of planar fractures that have a
length-to-height ratio greater than three.
− The model gives narrower widths and longer fracture lengths compared to
the KGD model.
− The excess pressure increases with times; in log-log coordinates, the slope
is positive.
− The model is most appropriate when the fracture length is much larger than
the fracture height.
The 2D fracture models require less input and perform the computations very
quickly. This advantage may be used when the fracture is known to be contained in
height, and the value of the height is known. However, the fracture shape is still
rectangular because no difference is calculated in the height at the tip of the fracture
compared to the height at the wellbore. Because of these conditions as well as the
1D fluid flow and noncoupling of the length and width, the 2D model fracture length is
usually longer, the net pressure is higher and the width is wider than actual or when
compared to the various P-3D or PL-3D models.

3.2 Pseudo Three-Dimensional (P-3D)


Most of the P-3D models incorporate basic assumptions regarding the elastic
properties of the rock layers, fluid flow and fracture initiation. These assumptions
reduce the complexity and number of calculations needed for a fracture simulation.
The elasticity is considered in two dimensions (but not in cross section) and fluid flow
is one-dimensional in the direction of the propagation. Many P-3D simulators
assume all layers have the same elastic, reservoir and fluid-loss properties. Some
P-3D simulators allow the user to input a description for each layer and then use
some method for averaging, while others use calculated shape factors to arrive at
the effect the layers will have on the geometry. For example, a different Young's
modulus could be input for 15 layers. The simulator, depending on the choice of the
author, may then use averaging or other factors to determine the actual impact over
the entire interval. The values of 15 different layers are now reduced to one. This
method can have significant implications in sections with thick layers of coal, shale or
other lithology. Most P-3D simulators have problems running when a large contrast
in elastic properties exists. The assumptions and use of averaging and factors allow
the P-3D simulators to generate a solution in a very short time, even on the small
personal computers.
The P-3D simulators divide the fracture into a fixed number of vertical elements in
the grid. Each vertical element extends from the top of the fracture to the bottom.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 16 of 35

All of the values describing the fracture geometry, pressure in the fracture and fluid
loss are the same throughout that one single element. This is one reason why most
P-3D simulators estimate more vertical height growth when compared to the PL-3D
simulators. Also, there is no pressure drop within the element from the perforations
to the vertical limits of the fracture. The pressure at the vertical tip of a P-3D fracture
simulation is also higher than that of a PL-3D simulation. The P-3D simulators may
produce solutions with excessive vertical height growth; however, the simpler
discretization scheme does significantly reduce the complexity and computational
time required to run a simulation.
Most P-3D simulators handle the fluid flow only in the horizontal direction within the
fracture. This fluid flow in the horizontal direction (the direction of the propagation) is
because of the single grid element in the vertical direction. This 1D flow also limits
the ability to properly describe the proppant transport within the fracture. To
compensate for the problem of proppant transport, some P-3D models use a
correlation that creates an elliptical proppant front.
P-3D simulations usually produce an elliptical fracture profile (side view) because of
the previously discussed methods for handling vertical growth, and because the P-
3D models are usually lumped models. Lumped models are those where the grid
data are integrated from the tip to the wellbore and averaged. This method typically
produces the elliptical fracture profile as illustrated in Fig. 7.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 17 of 35

Fig. 7. Elliptical profile (P-3D).

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 18 of 35

3.3 Planar Three-Dimensional (PL-3D)


PL-3D models are able to provide a detailed description of a hydraulic fracture in
both simple and very heterogeneous lithologic sections. These models require all of
the parameters describing the reservoir, elastic and fluid-loss properties for each
layer. Simulation can take 15 min or several hours depending on the complexity of
the problem and/or the computer being used.
PL-3D models divide the fracture into a mesh of vertical and horizontal grid
elements. The elements may take different shapes (usually small squares or
triangles). Fig. 8 shows an example of a mesh composed of quadrilateral elements
with triangular subelements.

Fig. 8. Example grid (PL-3D model).


A very fine mesh may be used around the edges and at the tip for greater accuracy,
while a coarser grid is used in the center of the fracture. The fracture is assumed to
develop as a plane (elasticity equations relate the pressure in the elements on the
crack faces to the width) and to calculate fracture-height growth. The different
values of pressure, fluid loss and other important parameters are calculated both
vertically and horizontally. The PL-3D model has 3D elasticity and 2D fluid flow.
Fracture (crack) propagation is controlled by the criterion of linear elastic fracture
mechanics. The fracture advances by a method where the stress intensity factor (KI)
is maintained nearly equal to the critical stress-intensity factor (KIc) during the crack
extension at each grid node. The theory of linear elasticity also allows problems to
be solved by the superposition of solutions to subproblems. Superposition is a
method whereby a PL-3D model can use part of Sneddon's solution to calculate the
vertical width profile.
Most PL-3D simulators allow the user to control the refinement of the grid. A coarse
grid with fewer elements will require less computing time, but there may be more
error in the solution. A very fine mesh grid will provide a solution with a very low
material balance error. The more complex grid system also allows the description of
the proppant transport in a more detailed manner.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 19 of 35

The 2D grid can produce some problems with complex stress profiles. Some PL-3D
simulators have difficulty handling bounding layers with a value of stress lower than
the pay, which are separated from the pay by a bounding layer with a higher stress.
This situation causes the grid elements to become increasingly skewed as the
fracture grows into the lower stress bounding layer. This continues until the
computational errors are extremely large and the simulator terminates the run.
The advantage of any PL-3D simulator is the ability to model the pressure drop
laterally within the fracture. As the fracture grows with each step of the simulation,
all of the parameters are recalculated in each grid element. Using the grid system,
the pressure at the lateral tip of the fracture can be lower than the pressure at the
wellbore. This allows the simulator to model a fracture with a greater vertical height
at the wellbore than at the fracture tip. The profile of this type of fracture is
determined by the grid and is not a lumped solution. Fig. 9 is a profile from a PL-3D
simulation.

Fig. 9. Fracture profile (PL-3D model).

3.4 Fully Three-Dimensional (3D)


Techniques for general 3D fracture propagation (including out-of-plane) have been
presented by some authors. The numerous problems and the system of equations
needed have not been developed to the point of providing a realistic, operational,
fully 3D model. A model that is truly a fully 3D model is needed to simulate special
conditions that cannot be handled by other models. However, at the present time,
no working model exists in the industry — except in research.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 20 of 35

4 Examples
The information used to compile the examples for the different models comes from a
case study where several calibration tests were performed. Extensive stress testing,
core analysis, reservoir evaluation and fluid analysis were also performed. The
reservoir parameters, stress values and other critical parameters are considered to
be the best available because of the many techniques used to obtain and
corroborate the data. Routine treatment designs usually never have such a large
volume of reliable data available with which to work.
Trying to examine all of the input and output data in detail for the examples, and for
each model, is simply too voluminous to cover in this section. Tables and profiles
will be used to show data and comparisons (case history). The actual model
comparisons will show profiles and an output summary for each model simulation.

4.1 Case History


This case history uses the same basic data that will be used later for the model
comparison. However, this case history uses 15 to 22 layers in the simulation,
whereas only 5 layers were used for the model comparisons given in Subsection 4.2
that follows this case history.
The reservoir stress and elastic properties that are used in the examples are given in
Fig. 10 and Fig. 11.

Fig. 10. Permeability, thickness and stress profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 21 of 35

Fig. 11. Computed values for Young's modulus and Poisson's ratio.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 22 of 35

Table 1 shows the depth, log stress and modified stress that can be used for layer
input. The modified stresses were obtained by modifying the log stresses to more
closely match the measured net pressure response from the calibration fractures.
These modified stress data were the values used for the model simulations.
Table 1. Comparison of Stress
Top of Zone Log Stress Modified Stress Delta Stress
(ft) (psi) (psi) (psi)
9030 7300 7300 0
9070 7800 8200 400
9115 7150 7350 200
9155 6600 6600 0
9170 6050 6050 0
9200 5600 5800 200
9250 5250 5250 0
9310 5850 6050 200
9340 6550 6550 0
9360 7300 7300 0
9380 5800 6200 400
9435 6400 6700 300
9455 7550 7950 400
9475 8400 8400 0
9575 7850 7850 0
Table 2 gives the permeability and fluid-loss coefficient used in the initial simulations.
Table 3 shows the design information for the treatment and the actual volumes used
during the treatment. The average injection rate was 50 bbl/min and a total of
1,168,910 lbm of sand was placed in the fracture. The average treating pressure
was 3000 psi.
Table 2. Permeability and Fluid Loss
Formation Permeability (md) 0.0065
Initial Total Leakoff Coefficient ( ft min ) 0.0010
Reservoir Fluid to Filtrate Permeability Ratio 10.0
Reservoir to Filter-Cake Permeability Ratio 100.0
Leakoff Interval Entire Fracture
The treatment was pumped down the casing/tubing annulus, and the bottomhole
pressure was measured by a pressure gauge run inside the tubing. Fig. 12 shows
the pressure profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 23 of 35

Note that the second calibration test performed on the well is also included on this
profile, and ends at a little over 100 min. A shut-in time is obtained and then the
main fracturing treatment starts at approximately 250 min. A PL-3D model
(GOHFER) from Marathon Oil Company was used to match the bottomhole pressure
as well as the surface casing (annulus) pressure. The simulation used stress and
elasticity data from the top of the log to the bottom. These data were used to give 22
layers for input. Fig. 13 shows the pressure match (dotted lines) from the start of the
treatment to the point of shutdown. Fig. 14 and Fig. 15 show the fracture length and
fracture width profiles from the simulation.

Table 3. Design Information


Fluid Volume, BBL Proppant Volume,
lbm
Stage Design Actual Fluid Type Proppant Design Actual Proppant
Conc. Type
(lbm/gal)
1 1500 1571 40-lbm linear 0 0 0 None
2 0 0 Shut-in 0 0 0 None
3 1000 576 Slickwater 0 0 0 None
4 3000 2908 50-lbm x-link 0 0 0 None
5 300 293 50-lbm x-link 1 12,600 11,110 100-Mesh
Sand
6 400 477 40-lbm x-link 0 0 0 None
7 400 337 40-lbm x-link 1 16,800 14,150 20/40
Ottawa sand
8 500 500 40-lbm x-link 2 42,000 41,120 20/40
Ottawa sand
9 500 437 40-lbm x-link 3 63,000 54,910 20/40
Ottawa sand
10 600 623 40-lbm x-link 4 100,800 105,330 20/40
Ottawa sand
11 800 755 40-lbm x-link 5 168,000 133,630 20/40
Ottawa sand
12 900 807 40-lbm x-link 6 226,800 227,250 20/40
Ottawa sand
13 900 934 40-lbm x-link 7 264,600 300,100 20/40
Ottawa sand
14 900 935 40-lbm x-link 8 302,400 292,420 20/40
Ottawa sand
15 300 240 Slickwater 0 0 0 None

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 24 of 35

Fig. 12. Profile of bottomhole, casing and tubing pressures.

Fig. 13. Pressure match for bottomhole and casing pressure.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 25 of 35

Fig. 14. Fracture profile.

Fig. 15. Fracture width profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 26 of 35

Net pressure was also matched using the P-3D Cleary model (FRACPRO) as run by
Resource Engineering Systems (RES). The net pressure match is shown in Fig. 16.

Fig. 16. Match of net pressure for calibration fracture and main fracture.
The drop in pressure at about 300 min, also shown on the bottomhole pressure plot,
corresponds to the time that the fracture height during the simulation reached a low
stress zone (see Stress Log Interval 9380 to 9455). Fig. 17 is the fracture profile
data from the simulation for the net pressure match.

Fig. 17. Fracture profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 27 of 35

Calculated Parameters
• Fracture Length (ft) 1819
• Propped Length (ft) 1441
• Fracture Height (ft) 358
• Propped Height (ft) 284
• Fracture Width (max.) (in.) 0.72
• Proppant Concentration (lbm/ft2) 1.88
These are the dimensions at the end of shut-in. Several methods were used to
attempt to determine the fracture height after the treatment. The method that was
accepted as being the most accurate in this case was the Continuous Microseismic
Radiation (CMR) log. This microseismic height log was run four months after the
treatment, and therefore will more accurately indicate the propped fracture height
rather than the created fracture height.
The CMR log indicated the height was from 9125 ft to 9375 ft, allowing ± 25 ft at both
top and bottom. This makes the propped height range from 250 ft to 300 ft. The
created fracture-height differences for the various models are difficult to evaluate
because created height can be significantly different from the propped height. This
difference can be caused by the way each model handles the information on the
various layers, and the actual width profile that was calculated. Another
consideration is how the model treats the proppant movement, settling or the many
other complicated aspects present when the proppant is added to the system.
Several postfracture reservoir evaluation techniques were used to analyze the
results of the treatment. The analysis testing was started after 89 days of production
from the well. A reservoir model was used to obtain the final history match of the
reservoir variables. The reservoir model used data obtained from buildup analysis
as well as the production match for both gas and water (two-phase flow). The result
of this analysis is illustrated in Fig. 18.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 28 of 35

Fig. 18. Reservoir model for final history match.


Note that conductivity was calculated to 1515 ft, but possibly only 1100 ft may be
truly effective. Total fracture half-length is shown as 1845 ft.
It must be pointed out that FRACPRO was the initial model used for the project, and
continued to be used through the completion of the project. Comparison with other
models (to be shown later) is more difficult because many times the final geometry
interpretation may be influenced by special data or variables input to obtain the
match. This is not wrong, and is the way that models should be calibrated.
However, other models may not have the advantage of using the information pieces
obtained throughout the time of the project. Also, in the case history just discussed,
the data given to run the simulation with GOHFER differed some from the exact input
used for FRACPRO.

4.2 Model Comparisons


This portion will show comparisons for a similar set of input parameters. The critical
output parameters obtained from simulating a common treatment (using the same
data as in Table 2 and Table 3) will be given, as well as profiles or graphs illustrating
the resulting fracture geometry. These comparisons were made by S.A. Holditch &
Associates, Inc., using original data from a project for the Gas Research Institute.
The reader will see some significant differences in some of the output values, but
these serve to show again the differences between the model code. Different
companies running the same model also obtain different results. This difference may
be caused by how each company handled the stress layers, and the fact that in
some cases the various companies were asked to model the calibration fracture and
use fluid-loss coefficients derived from that analysis. This left room for some to use a
wall-building coefficient initially, while others may have started with a total fluid-loss
coefficient. Hopes are that sometime in the future, one inclusive set of data will be
given to everybody and the models run again for comparison.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 29 of 35

The comparisons show results for KDG and PKN modes as well as for the P-3D and
PL-3D runs. Table 4 gives the list of output data from the runs using both GDK
(KGD)) and PKN geometry.
Note that since these models use fixed-height assumptions, most have the same
height listed because it is an input parameter. Comparisons are made for two cases
— (1) fluid viscosity set at 200 cp, and (2) viscosity based on n' and k' values. These
values give a viscosity of approximately 450 cp at 37.5 sec -1. Case 3 and Case 4
show the difference between fluid-loss values, which is also evident by the calculated
efficiency. For the sake of space, example profiles for these KGD and PKN
geometries will not be shown.
The list of output data from P-3D and PL-3D simulations is given in Table 5. These
simulations use the same data previously shown in Table 2 and Table 3. The same
information concerning the viscosities applies here as it did from the previous
discussion.
The additional data and comparisons for these P-3D and PL-3D simulations involve
the number of layers used in the particular simulation. The example profiles and
graphs for these simulations will be taken from Case 8 of Table 5 (Variable Viscosity,
5-layer). The detailed differences will not be discussed, but left to the reader to
examine.
Table 5 should be self-explanatory for comparing the information in Case 8 with the
figures. However, to eliminate confusion each simulation will carry the figure number
corresponding to the following list —
• Fig. 19. SAH (TRIFRAC) length and width profile.
• Fig. 20. NSI (STIMPLAN) length and width graphs (P-3D).
• Fig. 21. RES (FRACPRO) length and width profile (P-3D).
• Fig. 22. Marathon (GOHFER) length and width profile (PL-3D).
• Fig. 23. ARCO (TERRAFRAC) length profile only (PL-3D).
• Fig. 24. ARCO (STIMPLAN) length and width profile (P-3D).
• Fig. 25. MEYER length and width profile P-3D).
• Fig. 26. Ohio State length profile only (P-3D).
Copies of the simulation plots for length and width were not available for MEYER
(Bells) or TEXACO (FRACPRO).

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 30 of 35

Table 4. Fracture Model Comparison Runs

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 31 of 35

Table 5. Fracture Model Comparison Runs

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 32 of 35

Fig. 19. TRIFRAC length and width profile.

Fig. 20. STIMPLAN length and width.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 33 of 35

Fig. 21. FRACPRO length and width profile.

Fig. 22. GOHFER length and width profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture Modeling Dowell
Page 34 of 35

Fig. 23. TERRAFRAC length profile.

Fig. 24. STIMPLAN length and width profile.

DOWELL CONFIDENTIAL
Section 300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture Modeling
Page 35 of 35

Fig. 25. MEYER length and width profile.

Fig. 26. Ohio state length profile.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 1 of 52

Treatment Design
1 Introductory Summary............................................................................................................. 3
1.1 References........................................................................................................................... 3

2 Adaptive Design Methodology ............................................................................................... 6


2.1 Setting Characterization....................................................................................................... 7
2.2 Setting Modelization............................................................................................................. 8
2.3 Adaptive Treatment Design ............................................................................................... 10

3 Data Collection....................................................................................................................... 10

4 Fracturing Fluid Selection..................................................................................................... 12


4.1 Fracturing Fluid Selection Guide........................................................................................ 13
4.2 Water Sensitivity of the Reservoir Rock............................................................................. 16
4.3 Rheological Properties and Viscosity Requirements ......................................................... 17
4.4 Fluid Friction Pressure ....................................................................................................... 18
4.5 Fluid Compatibility with Reservoir Fluid and Rock ............................................................. 20
4.6 Rheology Selection ............................................................................................................ 20
4.7 Fracturing Fluid Additive Selection..................................................................................... 21

5 Proppant Selection ................................................................................................................ 21


5.1 Proppant Selection Methodology ....................................................................................... 22

6 FracCADE Software ............................................................................................................... 28


6.1 The FGS Module................................................................................................................ 29
6.2 The MLF Module................................................................................................................ 29
6.3 The FracNPV Module ........................................................................................................ 30
6.4 The INVERSE Module ....................................................................................................... 30
6.5 The PLACEMENT Module ................................................................................................. 31
6.5.1 PLACEMENT I........................................................................................................... 31
6.5.2 PLACEMENT II.......................................................................................................... 31
6.6 Additional FracCADE Modules........................................................................................... 31

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 2 of 52

7 Equations ................................................................................................................................33
7.1 Basic Equations..................................................................................................................33
7.2 Design Equations ...............................................................................................................34
7.3 Proppant Equations ............................................................................................................37

8 Proppant Flowback ................................................................................................................38

9 Refracturing ............................................................................................................................39
9.1 Candidate Selection ...........................................................................................................40
9.2 Refracturing Design Methodology ......................................................................................41
9.2.1 Estimation of Recoverable Reserves .........................................................................41
9.2.2 Estimation of the Present Fracture Geometry............................................................42
9.2.3 Determining the Average Reservoir Pressure ...........................................................44
9.2.4 Estimating Production Response From Refracturing .................................................44
9.3 The Effects of Fractured Well Pressure Distributions on Design........................................44
9.4 Effects of Reservoir Pressure Changes .............................................................................48
9.5 Refracture Treatment Design Considerations ....................................................................50
9.5.1 Fracture Initiation .......................................................................................................50
9.5.2 Fracture Extension.....................................................................................................50
9.5.3 Fracture Containment ................................................................................................52
9.5.4 Screenout...................................................................................................................52
9.5.5 Fracture-Fluid Recovery ............................................................................................52

FIGURES

Fig. 1. Basic fracture treatment design.........................................................................................4


Fig. 2. Ideal fracture treatment design..........................................................................................5
Fig. 3. Logic and associated activities of an adaptive procedure for fracture treatment design. .7
Fig. 4. The modelization concept relevant to treatment design. ...................................................9
Fig. 5. Fracturing fluid selection guide for a gas well..................................................................14
Fig. 6. Fracturing fluid selection guide for an oil well..................................................................15
Fig. 7. Effect on productivity index ratio (Jo, unfractured) of 5-in. damage around fracture........17
Fig. 8. Friction pressure drop of various tubing and casing sizes for 30 lbm/1000 gal delayed
(dashed lines) and non-delayed (solid lines) borate fluids. .............................................19
Fig. 9. Laminar and turbulent flow areas of viscoelastic fluids. ..................................................19
Fig. 10. Fracture half-length requirements for a gas well. ..........................................................23
Fig. 11. Retained permeability for linear fluids. ..........................................................................24
Fig. 12. Retained permeability for borate-crosslinked fluids.......................................................25

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 3 of 52

Fig. 13. Retained permeability for titanate-crosslinked fluids..................................................... 26


Fig. 14. Example of a production decline plot used to estimate recoverable reserves. ............. 42
Fig. 15. Example of a buildup data type-curve match. ............................................................... 43
Fig. 16. Example of a performance data type-curve match. ...................................................... 43
Fig. 17. Pressure contours around high-conductivity propped fractures.................................... 44
Fig. 18. The location of an infill well to improve recovery........................................................... 45
Fig. 19. Possible problem is fracture azimuth is unknown. ........................................................ 46
Fig. 20. Pressure contours around a short fracture or a fracture with low conductivity.............. 47
Fig. 21. Horizontal stress due to Poisson's ratio and pore pressure.......................................... 49
Fig. 22. Fracturing above a previous proppant-pack. ................................................................ 51

TABLES

Table 1. Core data example....................................................................................................... 12


Table 2. Rheology selection guidelines (pump time less than four hours)................................ 20
Table 3. Rheology selection guidelines (pump time greater than four hours)........................... 21
Table 4. Poppant Selection Guide ............................................................................................. 27

1 Introductory Summary
The purpose of hydraulic fracturing is to place an optimum fracture of certain
geometry and conductivity that will allow maximum incremental production (over that
of the unstimulated well) at the lowest cost. The objective in a treatment design is
either
• to achieve a fracture length and conductivity prescribed by the client
• to achieve an economic design, in terms of maximizing the net present value of
the treatment.
Basic fracture treatment design uses the methodology shown in Fig. 1.
Ideal fracture treatment design uses the methodology shown in Fig. 2.
Treatment design fundamentals for acid fracturing are provided in Acid Fracturing.

1.1 References
Thorough discussions of the fracturing process and parameters affecting the
fracturing process are provided in Reservoir Stimulation, A Practical Companion to
Reservoir Stimulation and other sections of this manual.
Computer-aided design information is provided in the FracCADE* User's Manual.
Additional information is provided in the SPE Monographs, Hydraulic Fracturing and
Recent Advances in Hydraulic Fracturing.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 4 of 52

Collect Formation and Well Formation

Look at Offset Well Information (If available)

Select Fluid

Select Proppant

Sufficient
Information (No) DataFRAC*
(Yes)

Select Model

Preliminary Pump Rate

Determine Net Pressure Limits Using


the FGS Module in the FracCADE Software

Determine the Most Cost-Effective Design Using


the FracNPV* Module in the FracCADE Software

Finalize the Pump Schedule Using


Pre-Pad the PLACEMENT Module in the
FracCADE Software

Surface Equipment Quality Assurance Pre-Treatment Job Flowback

Fig. 1. Basic fracture treatment design.


* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 5 of 52

Fig. 2. Ideal fracture treatment design.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 6 of 52

2 Adaptive Design Methodology


In terms of conventional materials science, rock masses are characterized by
nonstandard modes of behavior. These have resulted in the formation of a specific
methodology for the design of fracture treatments. The unorthodox response modes
include
• scale dependence of deformation and strength characteristics (that is,
constitutive behavior)
• the effect of load path on mechanical properties
• the difficulty of obtaining representative samples for determining mechanical
properties
• the effect of pore fluid on mechanical response
• the expression, in measured properties, of sample disturbance in the stress path
imposed on it during recovery and testing.
Applied to fracture engineering, the design methodology that accounts for these
modes of response is an adaptive design practice derived from well-established geo-
engineering procedures. It recognizes the need for recovering as much information
as possible from tests conducted on the rock mass in situ. In particular, for repeated
engineering activities such as a series of fracturing treatments in the same formation,
it seeks to recover the properties of the formation and its setting by analysis of the
observed response to previous treatments.
From the typical setting data provided by geophysical logs, the DataFRAC service or
a calibration treatment, a model of the field prototype is derived by idealization and
judgments about the state of the formation and setting. The model is necessarily
distorted relative to the true condition of the rock mass and this must be considered
when performing design exercises.
Logic
The logic of an adaptive design method for fracture treatment design is provided in
Fig. 3. The purposes of the method are:
• to account for the recognized deficiencies in reservoir characterization by
conventional exploration methods
• to recover information on the scale-dependent properties of the rock mass from
the full-scale loads imposed by fracture treatments
• to provide for progressive re-formulation of appropriate rock-mass and rock-
constitutive models.
The adaptive design method recognizes that on the reservoir scale, the quality of
fracture treatment design can be improved as the properties and in-situ state of the
rock mass are better defined by analysis of previously conducted fracture
treatments.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 7 of 52

The methodology provided in Fig. 3 consists of five phases, executed in a multi-pass


loop representing successive fracture treatments in a reservoir. It implies that, in
practice, data from the treatment of offset wells are to be used in the design of
treatments for adjacent wells.

Fig. 3. Logic and associated activities of an adaptive procedure for


fracture treatment design.

2.1 Setting Characterization


The aim of this phase is to define the mechanical and fluid conductive properties and
in-situ state of the medium in which hydraulic fracturing is to occur. Setting
characterization involves determination or estimation of the following.
• deformation
• shear (yield) strength
• fracture toughness
• porosity and permeability properties of the various lithological units in the
sequence, including the candidate formation or formations
• definition of the geometric and mechanical properties of natural fractures or
fracture sets in the candidate formations.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 8 of 52

An estimate of the in-situ shear strength may be made from relations between
properties such as estimated static Young's modulus and porosity. In evaluating the
in-situ state of stress, particularly the minor horizontal principal stress, the calculation
may be based on the depth stress and the assumption of uniaxial strain, requiring a
knowledge of the Poisson's ratio, which is usually derived from sonic logs.
The difficulty in formation characterization lies in having representative data defining
geomechanical conditions throughout the candidate setting. When formation
exploration is heavily based on well logs, limited physical access yields small
numbers of rock specimens (if any) for laboratory testing and imprecise theory exists
to relate specimen properties to those of the host medium. In these cases, a high
margin of uncertainty must be expected in the formation characterization
parameters*.

2.2 Setting Modelization


Modelization is a term conveniently applied to formulation of a setting model by the
simplification and rationalization of the data generated by setting characterization.
The aim is to account for the principal features that will control fracture and
deformation of the prototype. For example, after a layered model is developed from
log data, lithological units are ascribed dimensions and average “representative”
deformation properties, porosity and permeability. Assumptions are made about the
constitutive behavior of the rock in each layer. Natural fracture sets are assigned
regular geometry and shear-strength properties and a representative state of stress
is assigned to each layer in the sequence.
The need for elaboration of a setting model comes from two factors.
1. The limited data on in-situ conditions that can be collected by setting
characterization.
2. The limited detail that can be accommodated in all of the closed-form solutions
and analytical and numerical simulators that can be used in fracture design.
Some notions on the modelization concept relevant to treatment design are provided
in Fig. 4 which classifies modeling problems in terms of the amount or quality of data
defining the problem domain and the degree of understanding of the mechanics of
the system. All treatment design problems are initially “data-limited”, but the
mechanics of the process are arguably well understood; that is, the relevant domain
of the model definition map is zone 2. While the intention of an adaptive design
technique is to move the exercise into the “well-posed, well-understood” domain (that
is, zone 3), it also serves to emphasize the limitations to be accounted for at all
stages of a design exercise.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 9 of 52

Fig. 4. The modelization concept relevant to treatment design.

Because design exercises are conducted on a setting model that has assumptions
and simplifications embedded in it, the design engineer needs to assess the
sensitivity of design analyses to the limitations imposed by the model assumptions.
Explicit assumptions are made concerning factors such as thickness of layers,
homogeneity of properties and stress distribution. Frequently, the assumptions
about factors such as rock constitutive behavior and bonding between layers are
implicit, but nonetheless can render a design analysis meaningless if they are
incorrect. In particular, when numerical schemes are used in design analysis and
the results are questionable, the details of the setting model need to be examined for
mechanical consistency or validity. In contemporary geo-engineering practice, a
design utility frequently provides a capacity for interactive construction of a setting
model on which design analyses are then conducted, permitting ready identification
of invalid or inconsistent assumptions. Assuring model consistency with the
functionality of the design utility then becomes a routine but essential component of
setting modelization.
The most important point to note in relation to setting modelization is that the model
on which a design analysis is conducted is a distorted representation of the
prototype. The extent of the distortion is usually unknown, but techniques such as
Monte Carlo simulation of the possible range of setting parameters can be used to
identify the effects of distortions in the model.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 10 of 52

Most current fracture design practice does not recognize explicitly that a design
exercise is performed on a model of the field prototype. One result is that difficulties
in a design analysis are usually ascribed to some deficiency in the design utility (“the
code is useless”) rather than to a poorly conceived model of the setting. When
difficulties are encountered in a design analysis, it is frequently useful to consider if
the model is properly formulated relative to the functionality of the simulator, rather
than to conclude that the simulator is defective.

2.3 Adaptive Treatment Design


The treatment design is conducted in two steps. In the first step (the design
analysis), fracture fluid and proppant are selected, and a series of analyses is
conducted to determine the fluid injection rate and elapsed time (that is, the pump
schedule), which will meet the design objectives. In the second step (the treatment
simulation), the pump schedule is used in a fracture simulator to examine the
evolution of the fracture through the designed treatment and to calculate the length
and conductivity of the fracture generated by the treatment.
Exercise of judgment and several iterative cycles are implied in the design process.
For example, in the design analysis, identification of an optimum fracture length to
satisfy a net present value constraint implies iterating over a range of fracture
lengths. Further, if there is no closed-form expression relating fracture length, pump
rate and pump time, a method of successive approximations may be needed to
obtain a solution for the pump schedule. In the final treatment simulation phase, if
the discrepancy between the simulated fracture length and the design value exceeds
a reasonable engineering tolerance for geo-engineering design (perhaps 5%), it may
be necessary to repeat the design of the pump schedule. Some judgment may then
be required on the appropriate way to restart the design analysis cycle.

3 Data Collection
A thorough discussion of pretreatment-design data requirements is provided in
Reservoir Stimulation, Chapter 6. Unless the formation and setting data are
dependable and sufficient, a reliable fracture treatment design will not be possible.
In cases where the data set is insufficient, informed judgment may provide the
missing data, but the design should then be qualified by some assessment of the
effect on the design and the uncertainty in the data.
Log Interpretation Charts, Log Interpretation Principles/Applications, and Cased-Hole
Log Interpretation Principles/Applications from Schlumberger provides geophysical
and petrophysical logging information.
Client-Supplied Data
The client should provide administrative, well, reservoir, production, and rock-
mechanics data.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 11 of 52

Height Determination
Ideally, gross fracture height is determined using a FracHite* log (if calibrated with
actual stress data) in conjunction with the FGS or MLF module in the FracCADE
software.
Leakoff height (net reservoir height) can be based on a porosity cut-off or gamma/SP
deflection. Normally, the height of any zone with greater than 1/3 deflection from the
shale baseline is considered leakoff height.
Permeability Data
The only reliable source for formation permeability data is from pressure-buildup
analysis. Log-derived permeability and core-derived permeability can be unreliable.
Openhole log data can be used to estimate formation permeability. Virtually any
attempt to calculate reservoir permeability from a standard openhole logging suite
will use a relationship of the following form.

αφ b
k 1/ 2
= c
s wi
Where:
k = intrinsic permeability
φ = porosity
Swi = irreducible water saturation
a,b,c = constants.
Different forms of the equation are known as Tixier's equation, Timur's equation and
Coates' equation. These equations yield permeability estimates if effective
porosities, irreducible water saturations and bulk volume water saturations can be
determined from logs. Each of these implicitly assume that k = f (φ, Swi). This may or
may not be true. Rocks with intergranular porosity generally show a trend of
increasing permeability with increasing porosity and decreasing bulk volume water
saturation; however, permeability and porosity do not always scale in a tractable
fashion.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 12 of 52

Table 1 is an example of actual core data.

Table 1. Core data example.


Porosity (%) Bulk Volume Permeability (md)
Water Saturation
(%)
23.9 26.7 198
24.8 29.8 177
25.1 36.7 139
25.1 30.3 356
26.8 31.7 1105
26.9 33.1 1524
27.0 35.9 1790
27.9 36.5 632

Log-derived permeabilities can be reasonably good if sufficient work is done to


calibrate the analysis using actual core data, Repeat Formation Tester (RFT*) data
and well-test data. Studies using logs to differentiate between “high” and “low”
permeabilities in several different fields have been successful. Applying numbers to
“high” and “low” is sometimes difficult. For example, “high” may mean anything
greater than 100 md; however, in a zone with two high-permeability zones, one may
be 300 md, while the other may be 1000 md. Log-derived permeabilities (not
calibrated with other data) would not resolve the difference. Most clients will not
agree to a calibration step because it is time consuming, difficult and expensive. A
reliable fracture treatment design will not be possible using uncalibrated log-derived
permeabilities.
Do not rely on logs for permeability data in formations having vuggy porosity or
natural fractures or both.

4 Fracturing Fluid Selection


Fracturing fluids are required to efficiently open and propagate a fracture and
transport proppant throughout the length of the fracture. The selection of
appropriate fluids for a treatment has significant influence on the treatment
effectiveness and cost. Three considerations must govern fracturing fluid selection.
These are (in order of importance)
1. ability to create a fracture with large conductivity (that is, transport high slurry-
concentrations)
2. result in as little polymer-induced proppant-pack damage as possible
3. require lower pumping and treatment pressure capacity by reducing the friction

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 13 of 52

pressure drop.
All these considerations are affected by the polymer concentration that controls both
the fluid viscosity and the resulting friction pressure. Thus, the polymer
concentration must be engineered appropriately so that the concentration is
adequate but not excessive.
There are other less important considerations in fluid selection. While these should
be taken into account, they should not govern the fluid selection to the detriment of
the previously mentioned important concerns. These include minimization of
fracture-face damage, which would be the result of unavoidable leakoff, and
compatibility problems between the fracturing fluid and reservoir fluids and rock. In
addition, there has been much industry concern about post-treatment cleanup. This
concern has led to the use of energized and foam fluids. Although these fluids have
a decided edge on cleanup, energized and foam fluids become impractical when
super-high proppant concentrations are necessary because proppant is added
exclusively to the liquid portion of the fluid. The super-high proppant concentrations
may exceed the capabilities of the mixing and blending equipment. Foam fluids may
be more appropriate in very tight formations where fracture conductivity is less
important.
The fracturing fluid typically consists of two elements—the pad fluid and the slurry. A
prepad fluid is occasionally utilized.
Prepad Fluid
The prepad fluid is pumped first and is used to penetrate the matrix and initiate the
fracture.
Pad Fluid
The pad fluid follows the prepad fluid if a prepad fluid is pumped. If a prepad fluid is
not pumped, the pad fluid is pumped first and initiates the fracture. The entire
fracture is created by the pad fluid. Therefore, the pad fluid must be designed to
control filtrate-dependent leakoff and pressure-dependent leakoff to the formation.
Slurry
The slurry follows the pad and transports and places the proppant. The slurry
usually encounters only filtrate-dependent leakoff.

4.1 Fracturing Fluid Selection Guide


Fig. 5 is a fracturing fluid selection guide for a gas well. Fig. 6 is a fracturing fluid
selection guide for an oil well. While these selection guides should be considered as
general recommendations, the actual use (or even appropriate use) of fluids may be
lopsided toward certain types of fluids. Furthermore, these charts should always be
used with a degree of caution, bearing in mind the considerations previously
discussed. Fracturing fluid must be engineered with the particular reservoir in mind
and with consideration for the desired performance of the fracturing treatment.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 14 of 52

Fig. 5. Fracturing fluid selection guide for a gas well.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 15 of 52

Fig. 6. Fracturing fluid selection guide for an oil well.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 16 of 52

4.2 Water Sensitivity of the Reservoir Rock


The water sensitivity of the reservoir rock has traditionally prompted a division
between water-base fluids and oil-base fluids. However, this consideration is often
the cause of inappropriate fluid selection and less-than-optimum fracturing
treatments. Essentially, if the reservoir is mildly-to-moderately water sensitive, the
selection process outlined for a gas well should be followed.
While the use of oil-base fluids in oil wells is often suggested, these fracturing fluids
deserve certain additional considerations. The cost of pumping an oil-base fluid is
usually much greater than the cost of pumping a water-base fluid because of the
cost of the oil. Oil-base fluids, because of their inordinately high friction pressure
losses, typically exhibit high hydraulic horsepower requirements.
Safety considerations must also be addressed because of the flammability of the oil.
Any problem leading to fluid leakage has the potential for extreme fire hazard.
The perceived advantage of oil-base fluids is that the reservoir is exposed only to a
fluid that is related to the reservoir fluids. However, several studies have shown that
fluid leakoff and relative permeability-induced damage are not usually severe
problems. Fig. 7 shows that a damaged zone of 5 in. (much deeper than most fluids
will reach during leakoff) has a minimal effect on production as long as the fracture
has adequate conductivity. With this in mind, water-base fluids can be used in most
reservoirs without creating significant damage. Fig. 7 suggests that a tenfold
decrease in the reservoir permeability (or even a hundredfold decrease) has a very
limited impact on the productivity index ratio between the fractured and the
unfractured well. Fracture face damage should never be a criterion for the fracturing
fluid selection to the detriment of fracture conductivity.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 17 of 52

Fig. 7. Effect on productivity index ratio (Jo, unfractured) of 5-in. damage around fracture.

4.3 Rheological Properties and Viscosity Requirements


The fracturing fluid will almost totally suspend propping agents if the viscosity is
maintained at 100 cp ±25 cp at 170 sec-1 during pumping. When designing a fluid to
meet this requirement, the reservoir temperature must be known. The pad and half
of the slurry will be exposed to the bottomhole static temperature (BHST). The
fracturing fluid viscosity will decrease as a function of both increasing temperature
and increasing exposure time at this temperature.
Reservoirs with temperatures greater than 300°F (149°C) require fracturing fluids
with polymer concentrations of 40 to 60 lbm/1000 gal for the pad fluid and for the
early treating fluid stages. Polymer concentration may be reduced as the treatment
progresses because the later portion of the treatment will not be exposed to BHST
for an extended period of time. The polymer concentration in the final stages can be
as low as 20 lbm/1000 gal and the fluid may incorporate a less damaging crosslinker
(that is, borate instead of an organometallic material). Precision continuous mixing
allows the engineering and tailoring of fracture fluids and additives to yield optimum
properties.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 18 of 52

Waterfrac Fluids
Waterfrac (WF) fluids can be successfully applied in thin, low-temperature reservoirs
(less than 150°F [66°C]) with excellent barriers. Normally, WF fluids are pumped at
0.5 to 1 bbl/min per foot of gross height. The maximum proppant concentration is
typically less than five lbm proppant added. In this case, proppant continuously
settles in the fracture and creates proppant banking.

4.4 Fluid Friction Pressure


Pumping highly non-Newtonian viscoelastic fluid through the treating tubulars results
in significant friction pressure. This pipe friction pressure cannot be accurately
calculated using the classic Reynolds number/friction-pressure methods; however
the pipe friction pressure can be calculated using methodology provided in
Treatment Execution.
Fluid friction pressure in the tubulars, through the perforations, and along the
fracture has a direct relationship with fluid viscosity. Fluid friction pressure through
the perforations and along the fracture are usually low and are therefore neglected.
The exception to this is in limited entry (discussed in Diverting Techniques).
Fig. 8 provides friction pressure guidelines for Widefrac 130 (YF130) fluids for
various tubing and casing sizes. The dashed lines are for completely delayed
crosslinked fluids, and the solid lines are for nondelayed fluids. The observed break
in the curves corresponds to the transition from laminar to turbulent flow (Fig. 9).
For example, if a YF130 fluid is pumped at 20 bbl/min through 2.441 in. tubing, the
friction pressure is equal to approximately 320 psig/1000 ft.
A polymer concentration increase will cause a friction pressure increase. Large
differences will be observed at low flow rates in small-diameter tubing.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 19 of 52

Fig. 8. Friction pressure drop of various tubing and casing sizes for 30 lbm/1000 gal
delayed (dashed lines) and non-delayed (solid lines) borate fluids.

(This Figure is for illustration purposes only.)

Fig. 9. Laminar and turbulent flow areas of viscoelastic fluids.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 20 of 52

4.5 Fluid Compatibility with Reservoir Fluid and Rock


Chemical analysis of the reservoir fluid and rock may be considered when selecting
a fracturing fluid. Clay content, iron precipitation, asphaltene and paraffin deposition,
are among the compatibility considerations.

4.6 Rheology Selection


The selection of fracturing fluid rheology for design purposes may be determined
using the following method.
1. Estimate the pad volume, slurry volume, and pumping time using the FGS or
MLF module in the FracCADE software. For estimation purposes, use 0.6 as a
value for n’ and 0.02 as a value for K’ in the fluid editor. The maximum pump
rate should be the pump rate at 75% to 80% of maximum allowable surface
pressure.
2. Select fluid rheology using the guidelines provided in Table 2 for pump times
less than four hours. Use the guidelines in Table 3 when pump time exceeds
four hours. Fluids 1 to 3 (and 4, if applicable) are selected at the exposure
temperature and time to maintain 100 cp ±25 cp at 170 sec-1. The exposure time
for fluid 3 (and 4, if applicable) is based on the BHST; for a BHST near 150°F
(66°C), the exposure temperature would be 80% of the BHST.
3. If the treatment volume is small (less than 40,000 gal), one or two fluids may be
sufficient.

Table 2. Rheology selection guidelines


(pump time less than four hours)
Fluid Fluid Exposure Temp, (°°F [°°C]) % of Pump Time
Index
pad 1 BHST 50
½ slurry 2 BHST 50
½ slurry 3 70% of BHST 25
(200 to 300
[93 to 149])
80% of BHST
(150 [66])
90% of BHST
(100 [38])

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 21 of 52

Table 3. Rheology selection guidelines


(pump time greater than four hours)
Fluid Fluid Exposure Temp. (°°F [°°C]) % of Pump Time
Index
½ pad 1 BHST 35
½ pad 2 BHST 65
½ slurry 3 BHST 50
½ slurry 4 70% of BHST 25
(200 to 300
[93 to 149])
80% of BHST
(150 [66])
90% of BHST
(100 [38])

4.7 Fracturing Fluid Additive Selection


Appendix C  Additives provides guides for selecting fracturing fluid additives.

5 Proppant Selection
Adequate fracture flow capacity and ideal placement of the propping agent are
considerations vital to the design of a proppant program.
The more important parameters affecting fracture flow capacity are
• the physical properties of the propping agent
• proppant concentration in the fracture
• closure stress
• fracture width after closure
• contaminants (for example, fracturing fluid residue).

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 22 of 52

5.1 Proppant Selection Methodology


Optimum proppant selection uses the following methodology.
1. Calculate the required in-situ fracture permeability using Eq. 1.

100 C fD keff x f
kf = (1)
w (%retained permeability )
Where:
kf = in-situ fracture permeability (md)
CfD = fracture conductivity (dimensionless),
keff = formation effective permeability (md)
xf = fracture half-length (ft)
w = fracture width (in.).
If the formation effective permeability is less than 0.1 md, use a value of 10 for the
fracture conductivity. If the formation effective permeability is greater than 0.1 md,
use a value of 5 for the fracture conductivity.
The optimum value for fracture half-length can be determined using the FracNPV
module in the FracCADE software. A value for the desired fracture half-length may
also be obtained from the client or may be estimated (for a gas well) from Fig. 10.
A value of 0.1 to 0.15 in. is valid for fracture width in a PKN fracture. A value of 0.2
to 0.3 in. is valid for fracture width in a KGD fracture.
The retained permeability is determined using Fig. 11 (linear fluids), Fig. 12 (borate-
crosslinked fluids) or Fig. 13 (titanate-crosslinked fluids). Retained permeability for
zirconate-crosslinked fluids, oil-base fluids, and emulsion-base fluids will be included
in future revisions of this document as the data becomes available. Until then,
retained permeability may be determined in the laboratory as required.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 23 of 52

Fig. 10. Fracture half-length requirements for a gas well.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 24 of 52

Fig. 11. Retained permeability for linear fluids.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 25 of 52

Fig. 12. Retained permeability for borate-crosslinked fluids.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 26 of 52

Fig. 13. Retained permeability for titanate-crosslinked fluids.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 27 of 52

2. Select proppants from Table 4. Input data for the selected proppants into the
proppant editor module in the FracCADE software. Long-term permeability data
for propping agents is provided in the FRACTURING MATERIALS MANUAL 
ADDITIVES and the proppant database in the FracCADE software.

Closure Table 4.
Stress Poppant Selection Guide
(psi)
Required In-Situ Fracture Permeability (Darcies)
<<100 101 to 200 201 to 300 301 to 400 401 to 600 601 to 800 801 to 1000 >1000
<2,000 20/40 NWS 16/30 NWS 12/20 NWS 12/20 TBS
20/40 TBS 16/30 TBS
3,000 20/40 NWS 16/30 TBS 16/30 & 12/20 12/20 ISP
NWS
20/40 TBS 12/20 TBS
4,000 20/40 NWS 16/30 & 12/20 12/20 CRCS 20/40 ISP 16/20 ISRLD 16/20 ISP
NWS
20/40 TBS 16/30 & 12/20 16/30 PCRCS
TBS
20/40 CRCS 16/30 CRCS 20/40 ISP-LD
20/40 PCRCS
5,000 All NWS 16/30 TBS 16/30 PCRCS 16/20 HSRCS 16/20 ISP-LD 16/20 ISP 12/20 ISP
20/40 & 12/20 20/40 CRCS 20/40 ISP-LD 20/40 ISP 20/40 HSP
TBS
20/40 PCRCS 20/40 HSRCS 20/40 HSP
20/40 HSRCS 12/20 HSRCS
6,000 All TBS 20/40 PCRCS 16/30 PCRCS 16/20 ISP-LD 12/20 ISP
All CRCS 20/40 HSRCS 16/30 HSRCS 20/40 ISP 16/20 ISP
20/40 ISRLD 20/40 HSP
7,000 20/40 PCRCS 16/30 PCRCS 16/20 ISP-LD 12/20 & 16/20 20/40 HSP
ISP
12/20 HSRCS 16/30 HSRCS 20/40 ISP
20/40 ISP-LD
20/40 HSRCS
12/20 HSRCS
8,000 20/40 PCRCS 16/30 PCRCS 16/20 ISP 20/40 HSP
20/40 HSRCS 16/30 HSRCS
20/40 ISPLD
20/40 HSRCS
16/20 ISP-LD
20/40 ISP-LD
20/40 ISP
9,000 20/40 HSRCS 20/40 & 20/40 HSP
120/20 ISP
16/30 HSRCS
>10,000 20/40 & 12/20 ISP 16/20 ISP 20/40 HSP
20/40 HSRCS

NWS Northern White Sand


TBS Texas Brown Sand
CRCS Curable Resin-Coated Sand
PCRCS Precured Resin-Coated Sand
HSRCS High-Strength Resin-Coated Sand
ISP Intermediate-Strength Proppant
ISP-LD Intermediate-Strength Proppant (low density)
HSP High-Strength Proppant

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 28 of 52

6 FracCADE Software
The FracCADE software is used to design and evaluate the fracturing treatment.
The FracCADE software simulates fracture geometry, proppant placement and
distribution, and well performance. User information for the software is provided in
the FracCADE User's Manual.
The following sections present brief descriptions of FracCADE software utilities in the
context of an adaptive design methodology for fracture engineering. While the
application of these utilities appears straightforward, it is useful to note some
important issues that can affect their successful, effective and reliable application in
engineering operations.
Treatment Design and Evaluation Utilities
One set of utilities (MLF, INVERSE-NUMERICAL and PLACEMENT II) provides a
coherent and internally consistent scheme for fracture design and evaluation,
independent of other utilities. In particular, they may be applied as a set exclusive of
the current fracture design and simulation applications (FGS, INVERSE-
ANALYTICAL and PLACEMENT I). The physics and mechanics embedded in MLF,
INVERSE-NUMERICAL and PLACEMENT II are a better representation of rock
mass and fracturing fluid behavior under fracturing conditions than those in the
current codes. Therefore, detailed correspondence is not to be expected between
fracture designs conducted with FGS, INVERSE-ANALYTICAL and PLACEMENT I
and MLF, INVERSE-NUMERICAL and PLACEMENT II, except in the unlikely
circumstance where there is exact consistency between both the design
assumptions and the governing mechanics for the particular case.
Importance of Accurate Data
In the application of the utilities, the increased attention to the details of the
mechanics of fracturing expressed in the codes requires that this level of detail be
properly defined in the data set provided for a fracture design. This means that more
input data is required to design a treatment with the MLF, INVERSE-NUMERICAL
and PLACEMENT II utilities. Unless the formation and setting data are dependable
and sufficient, a reliable fracture treatment design will not be possible. In cases
where the data set is insufficient, informed judgment may provide the missing data,
but the design should then be qualified by some assessment of the effect on the
design and the uncertainty in the data.
Formation Model
Setting characterization data is used to construct a model of a formation, on which a
design exercise is performed. The model is a distorted representation of the field
prototype. It is imperative that the design engineer keeps in mind that the distortion
that is embedded in the setting model determines both the reliability of a treatment
design and the execution success of the fracture treatment. Further, in the
formulation of the model of the formation and setting, unless the model is consistent
with the functionality of the design utilities, it is impossible to perform the design

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 29 of 52

analysis in a logical and technically sound way. Because all design utilities are in
some way limited in functionality, one role of the design engineer is to understand
the functionality of any design utility in sufficient detail to permit formulation of a
compatible setting model.
Rock Behavior
Some substantial simplifications have been made in representing the constitutive
behavior of rock. As implied by the slopes of log-log plots (bottomhole treating
pressure versus time), the conventional representation of rock as an isotropic, linear
elastic continuum subject to brittle fracture is tenable for a wide range of fracture
treatments. However, there are formation conditions in which more complex
constitutive behavior are expressed. These include high-porosity, low-cohesion
formations that may exhibit pore consolidation and plastic yield under fracturing
conditions, and heavily fractured formations that may yield at low deviatoric states of
stress. Although these conditions may be recognized from pressure-history plots or
from the low fracture compliance expressed after shut-in, they are not properly
accounted for in MLF, INVERSE-NUMERICAL and PLACEMENT II (or in any other
commercial fracture-design utilities).

6.1 The FGS Module


The FGS module in the FracCADE software is used to perform sensitivity studies for
horizontal, KGD, PKN and radial fracture geometries. The module will calculate the
fracture geometry based on the total fluid volume pumped. The module will also
inversely calculate the injection fluid volume based on the desired geometry.
Analytical computations take only a few seconds which makes the module attractive.
Seven variables are identified as sensitivity parameters. These are:
1. fluid type
2. fracture height
3. leakoff coefficient
4. Young's Modulus
5. pump rate
6. spurt loss
7. rock toughness.
The calculated net pressure can be used in conjunction with a FracHite log (if
calibrated with actual stress data) to determine fluid viscosity, pump rate, and fluid
volume limitations.

6.2 The MLF Module


The MLF module simulates the simultaneous initiation and extension of multiple
hydraulic fractures. It couples a single fracture model to a set of constraints which

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 30 of 52

represent the hydraulic coupling of the fractures within the well. Significant features
of the model are:
• transient fluid flow is permitted
• proppant transport is tracked
• screenouts are modeled
• flow reversal is permitted.
Currently, MLF supports PKN fractures only.

6.3 The FracNPV Module


The FracNPV module is used to calculate the most cost-effective fracture design.
Production forecasting, fracture geometry and economic analysis are coupled in the
FracNPV module. Well, reservoir, fluid, and proppant data and operational
constraints, fixed costs and production constraints are used to calculate the optimum
fracture design.
Parametric studies can be performed for the following variables.
• fracture fluid
• proppant type
• proppant concentration
• fracture height
• leakoff coefficient
• formation permeability
• retained permeability
• pump rate.
Values obtained from the FracNPV output can be used in the PLACEMENT or
INVERSE module to design the optimum fracture treatment by maximizing economic
returns and minimizing treatment costs.

6.4 The INVERSE Module


The INVERSE module is used to develop a pump schedule from a given set of
parameters resulting in a fracture with the desired propped length and fracture
conductivity. The module allows input of a maximum (surface) proppant
concentration or the desired propped fracture concentration after closure.
The desired fracture length, proppant concentrations, mechanical constraints and
fracture-height constraints should first be determined using the FracNPV module.
The optimum values or the client requirements should then be simulated using the
INVERSE module.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 31 of 52

INVERSE  ANALYTICAL
The analytical simulator is used in conjunction with the FGS and PLACEMENT I
modules.
INVERSE  NUMERICAL
The numerical simulator is used in conjunction with the MLF and PLACEMENT II
modules.

6.5 The PLACEMENT Module


The PLACEMENT module is used to optimize the treatment design (fracturing fluid,
proppant and the pump schedule).

6.5.1 PLACEMENT I
PLACEMENT I is a two-dimensional (2D) hydraulic fracture simulator that provides
the capability to model fixed height fracture extension. The simulator will not
continue after proppant bridging or fluid dehydration. The fracture geometries
supported by PLACEMENT I are horizontal, KGD and PKN.

6.5.2 PLACEMENT II
PLACEMENT II is a pseudo three-dimensional (P3D) hydraulic fracture simulator
that provides the capability to model fracture growth into layers above and below the
pay zone along with fracture growth and recession. By modeling extension and
recession, tip screenouts can be simulated. The simulator can continue after
proppant bridging or fluid dehydration. A lateral coupling option is available which
represents the gradual evolution of a fracture from a short KGD-type fracture to a
longer PKN-type fracture. The fracture geometries supported by PLACEMENT II are
horizontal, KGD, P3D, P3D lateral coupling, PKN, PKN lateral coupling and radial.

6.6 Additional FracCADE Modules


JOB RECORD DATA ENTRY Module
The JOB RECORD DATA ENTRY module allows ACSII and PPR* datafiles to be
read into the FracCADE interface for the PRESSURE ANALYSIS, DECLINE DATA
and DataFRAC modules.
PRESSURE ANALYSIS Module
The PRESSURE ANALYSIS module is used to extract portions of the main datafile
for each of the different tests performed during the DataFRAC service and to
generate a graphical representation of the job data.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 32 of 52

DECLINE DATA Module


The DECLINE DATA module is used to calculate the hydrostatic pressure change
resulting from wellbore fluid heating during the DataFRAC calibration service. After
shut-in, the wellbore fluid temperature increases, decreasing the fluid density and
therefore, the hydrostatic pressure. If the bottomhole pressure is calculated using
surface pressure measurements, this hydrostatic pressure change must be
considered. The wellbore fluid temperature calculation model is a numerical finite
difference scheme using the superposition approach.
DataFRAC Module
The DataFRAC module is used to analyze the post-fracture pressure decline and to
perform the calculations necessary for aid in the pressure-decline analysis. The
parameters that can be determined are
• closure pressure
• the appropriate 2D model
• fluid efficiency (fluid used in the DataFRAC service
• fracture geometry at the end of the DataFRAC service)
• effective total leakoff coefficient.
QUICK Module
The QUICK module is used to generate a preliminary fracturing proposal in less than
45 minutes for a range of permeabilities and producing pressures. The QUICK
module calculates the fluid and proppant volumes required for a given fracture
length. The fluid volume calculations are based on fluid efficiency. The proppant
volume calculations are based on the fracture volume and maximum proppant
concentration. The pumping schedule is calculated from the slurry volume and
proppant concentration profile.
TUBING Module
The TUBING module is used to calculate the applied, pressure-induced and
thermally-induced forces on the tubing during a stimulation treatment.
FOAM SCHEDULE Module
The FOAM SCHEDULE module is used to calculate the liquid, gas, foamer and
proppant amounts. The FOAM SCHEDULE module is linked to the PLACEMENT
and INVERSE modules. The module allows:
• commingled gases
• foamer injection at different points
• options for foam quality compensation due to proppant addition
• the maximum proppant concentration (at the blender) can be used to define the
schedule.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 33 of 52

ACID SIMULATOR Module


The ACID SIMULATOR module allows a comprehensive simulation of different acid
reaction mechanisms. The PLACEMENT II simulator includes reaction-kinetics
models coupled with different fracture geometry models. The module will simulate
pad-acid-flush treatments, DUOFRAC* II treatments, acid-flush treatments, gelled
acid systems, LCA* systems and retarded acid systems.
BRACKETFRAC Module
The BRACKETFRAC* module is used to design fracture-height containment
treatments. The module determines the distance from the wellbore where the net
pressure becomes less than the maximum allowable net pressure governed by the
integrity of the barriers. This distance is the penetration required for coverage by
diverting agents.
FORECAST Module
The FORECAST module is a two-dimensional, single-phase reservoir simulator.

7 Equations

7.1 Basic Equations


Fracture Gradient (Eq. 2)

Pisi ρ f D
Pfg =
D (2)
Bottomhole Fracturing Pressure (Eq. 3)

Piw = pw + ph − pipe − Ppf (3)


Initial Shut-in Pressure (Eq. 4)

Pisi = piw − ph = Pfg − ph = pc + ∆ pf


D
(4)
Surface Fracturing Pressure (Eq. 5)

Pw = piw + p pipe + p pf − ph (5)


Hydraulic Horsepower (Eq. 6)

HHP = 0.0245 pw q (6)


Where:

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 34 of 52

D = depth (ft)
HHP = hydraulic horsepower
pc = closure pressure (psig)
pfg = fracture gradient (psig/ft)
ph = hydrostatic pressure (psig)
pisi = initial shut-in pressure (psig)
piw = reservoir fracturing pressure (psig)
ppf = perforation friction pressure (psig)
ppipe = tubular friction pressure (psig)
pw = surface fracturing pressure (psig)
q = pump rate (bbl/min)
∆pf = net pressure (psig)
ρf = density of the fracturing fluid (lbm/gal).

7.2 Design Equations


Shear Modulus (Eq. 7)

E
G = 0.5
1− v (7)
Closure Pressure (Eq. 8)

Pc = piw − pwf (8)


Dimensionless Fracture Conductivity (Eq. 9)

kfw
C fD =
kx f (9)
Proppant Settling (Stoke's Law) (Eq. 10)

1/ n'
( 2n' +1)d p  (ρ p − ρ f )d p 
Vs =
108 n'  72K ' 

(10)
Reservoir Fracturing Pressure (Eq. 11)

Piw = pc + ∆pf (11)


Material Balance During Pumping (Eq. 12)

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 35 of 52

qi t p = 2wx f h + 3Cδ f A f t p (12)


Fracture Compliance (Eq. 13)

(13)
w = c f ( piw − pc ) = c f ∆pf
Fluid Efficiency (Eq. 14)
v fp
η= (14)
Vi
Fracture Width During Pumping (PKN) (Eq. 15)

1/ 4
 µqx f 
w=  (15)
 E 

Fracture Width During Pumping (KGD) (Eq. 16)

1/ 4
 µqx f 
2

w =   (16)
 Eh 
Propped Fracture Length (Eq. 17)

.0605 Wp
Lf =
wh (17)
Apparent Viscosity (Eq. 18)

47880 K '
µa =
y 1− n' (18)
Friction due to Proppant (Eq. 19)

Pf , slurry = f m ( Pf , base )
(19)
To use Eq. 19, the proppant friction coefficient (sfexp) must be calculated. The
proppant friction coefficient varies with proppant size, fluid type, pipe size and well
deviation. For similar treatments in vertical wells, sfexp should be consistent. The
default value for sfexp in the PPR software is 0.35; however, field locations have
reported values up to 1.2. The proppant friction coefficient can be calculated using
methods provided in TREATMENT EXECUTION.
Where:

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 36 of 52

Af = area of one fracture face (cm2)


cf = fracture compliance
C = fluid-loss coefficient (ft/min½ )
CfD = fracture conductivity (dimensionless)
dp = proppant diameter (in.)
E = modulus of elasticity, Young's Modulus (psig)
fm = friction multiplier (dimensionless)
G = shear modulus (psig)
h = gross height (ft)
k = permeability (md)
kf = in-situ fracture permeability (md)
K′ = consistency coefficient (lbf-secn’/ft2)
Lf = fracture length (ft)
n′ = power-law exponent (dimensionless)
pc = closure pressure (psig)
pf,base = frictional pressure drop of base fluid through tubulars (psig)
pf,slurry = frictional pressure drop of slurry through tubulars (psig)
piw = reservoir fracturing pressure (psig)
pwf = bottomhole flowing pressure (psig)
q = pump rate (bbl/min)
qi = constant pump rate (bbl/min)
tp = total pump time (min)
Vfp = fracture volume (ft3)
Vi = volume injected (ft3)
Vs = terminal settling velocity (ft/sec)
w = fracture width (in.)
Wp = proppant weight (lbm)
xf = fracture half-length (ft)
γ = shear rate (sec-1)
δf = ratio of fluid-loss area to fracture area
η = fluid efficiency (dimensionless)
µ = fracture fluid viscosity (cp)

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 37 of 52

µa = apparent viscosity (cp)


v = Poisson's ratio
ρf = density of the fracturing fluid (g/cm3)
ρp = proppant absolute density (lbm/ft3)
∆pf = net pressure (psig).

7.3 Proppant Equations


The slurry density weight, specific gravity of the slurry or liquid or proppant volume in
the slurry can be calculated using the following equations.
Pounds of Proppant in One Gallon of Slurry (Eq. 20)

x
x
1+
8.345 SG p (20)
Weight Density of Slurry (lbm/gal) (Eq. 21)

x + 8.345 SGL
x
1+
8.345SG p (21)
Specific Gravity of Slurry (Eq. 22)

x + 8.345SGL
x
8.345 +
SG p (22)
Gallons of Liquid in One Gallon of Slurry (Eq. 23)

x
1−
8.345SG p + x (23)
Slurry Yield (Eq. 24)

x
1+ (24)
8.345 SG p

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 38 of 52

Where:
SGL = specific gravity of the liquid phase
SGp = specific gravity of the proppant
x = pounds of proppant added to a gallon of liquid (lbm).

8 Proppant Flowback
Proppant flowback is affected by the following.
• closure stress
• fracture aperture
• rock hardness/proppant embedment
• proppant type
• proppant size
• particle distribution
• existence of channels (incomplete fracture fill)
• pressure drawdown.
Pressure drawdown is probably a major cause for early-time proppant flowback
during the cleanup phase due to increases in viscous effects (multiphase flows) and
non-darcy effects (flow velocities are generally high and beta can be in the order of
four to five times higher than for dry gas or single-phase flow conditions).
The following papers discuss the problem of proppant flowback and offer some
insight into the mechanisms and what can be done to minimize the problem.
1. Milton-Taylor, D., Stephenson, C., Asgian, M., SPE 24821 Factors Affecting the
Stability of Proppant in Propped Fractures: Results of a Laboratory Study, 67th
Annual Conference, Washington DC, October, 1992.
2. Norman, L.R., Terracina, J.M., McCabe, M.A., Nguyen, P.D., SPE 20640
Application of Curable Resin-Coated Proppants, 65th Annual Conference, New
Orleans, LA, September 1990.
3. Martins, J.P., Abel, J.C., Dyke, C., Michel, C., Stewart, G., SPE 24858 Deviated
Well Fracturing and Proppant Production Control in the Prudhoe Bay Field, 67th
Annual Conference, Washington DC, October, 1992.
4. Martins, J.P., Milton-Taylor, D., Leung, H.K., SPE 20709 The Effects of Non-
Darcy Flow in Propped Hydraulic Fractures, 65th Annual Conference, New
Orleans, LA, September, 1990.
5. Maloney, D.R., Gall, B.L., Raible, C.J., SPE 16899 Non-Darcy Gas Flow
Through Propped Fractures: Effects of Partial Saturation, Gel Damage and
Stress, 62nd Annual Conference, Dallas, TX, September, 1987.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 39 of 52

Unfortunately, much like sand control, the prediction of proppant flowback is


site/field/formation/fracture specific and dependent on field observation correlations
developed for the area.
To minimize proppant flowback, consider the following:
• screenout designs (maximum packing)
• minimize proppant settling during placement and closure periods
• allow longer term shut-in periods (overnight) prior to initiating cleanup (for
maximum viscosity reduction and proppant-packing rearrangement)
• restrict initial flowback rates during the cleanup phase
• enlarge the perforated interval to reduce the wellbore flow velocities.
A mechanical screen, the STIMPAC* Service or resin-coated proppants can be
considered as a last resort.

9 Refracturing
Refracturing has a long and successful history. A review of well data from the 1960s
shows that approximately 35% of the fracturing treatments were refracturing
treatments. The fact that small volumes of proppant were used in the early
treatments made most of the wells good candidates for refracturing. More recently,
wells with massive hydraulic fractures have been successfully refractured. The
engineering considerations for refracturing treatments are generally the same as
those for any fracturing treatment. Some exceptions are
• More information is usually available for the design of the refracture treatment
(for example, previous well treatment data) and there is often more history of
fracturing in the area than was available for the initial treatment.
• Reservoir conditions may have changed since the initial treatment. The reservoir
pressure is usually less due to the volume of production. There can be significant
horizontal pressure gradients in the reservoir due to production.
• There is probability that the proppant from the previous treatment will act as
diverting material during the refracturing treatment. This may or may not
increase chances of retreatment success.
Refracturing treatment success is approximately 75% and is wholly or partially
attributed to the exceptions listed above. Contributing significantly to the success
are the subsequent technology advances in fluids, quality and fracturing techniques
that have been made since the initial fracturing treatment.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 40 of 52

Refracturing Economics
The economics of refracturing is based on reserve estimates, rate of recovery and oil
or gas price. The reserves in gas wells are dependent upon reservoir pressure,
formation thickness and porosity. The rate of recovery is determined by reservoir
pressure, formation thickness, porosity, permeability, fracture length and fracture
conductivity.
Infill Drilling
Infill drilling is a competing operation to refracturing. The advantage to infill drilling is
that an area of greater reservoir pressure can be penetrated directly by the new
wellbore if properly located. Refracturing penetrates the lower pressure reservoir
area near the wellbore prior to reaching an area of greater pressure. Some success
of infill drilling can be attributed to penetrating previously isolated portions in the
reservoir.

9.1 Candidate Selection


A candidate well for refracturing is
• a well that received a comparably small initial fracturing treatment by design
• a well in which the initial fracturing treatment was terminated early because of a
screenout or mechanical problems
• a well which has seen a significant production decline that is attributed to the loss
of propped fracture conductivity (for example, proppant crushing, plugging, or
scaling)
• a well that had poorer response from the initial fracturing treatment when
compared to other wells in the field
• a well containing multilayered intervals in which some intervals may not have
been fractured.
The client will normally pick the candidate for refracturing based on estimated
reserves, projection of the present production rate, success of other fracture
treatments in the field, available new technology and some comparison of the cost vs
additional recovery. Additional costs and risks other than those of the refracturing
operation often make the prospect unattractive. Fracturing technology changes
during the development of a field can improve results, making the early completions
in the field candidates for refracturing.
All candidates require adequate remaining reserves in order to warrant the
investment of the workover. A good estimate of reservoir parameters (that is,
permeability, porosity, net height and average reservoir pressure) is important.
Since the formation permeability is usually not known, the decision for refracturing
will be based on a comparison of production rates of the well with others in the field,
assuming all the wells have the same permeability and reservoir pressure. Poor
production response after refracturing can often be attributed to poorer than

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 41 of 52

expected reservoir properties. This can be minimized to some extent by only


considering wells that had good initial response from the original fracturing
treatment.

9.2 Refracturing Design Methodology


The following methodology can be used for refracturing treatment design.
1. Estimate the hydrocarbon reserves using current data.
2. Simulate the initial fracture treatment using the FracCADE software.
3. Evaluate well-test or production data (or both) to determine the state of the
present fracture and resolve any differences using the FracCADE software.
4. Extrapolate production using current well conditions.
5. Make a production forecast using the proposed refracture dimensions.
6. Make an economic comparison using step 4 and step 5.
7. Design the refracturing treatment using the FracCADE software.

9.2.1 Estimation of Recoverable Reserves


Recoverable reserves are estimated by extrapolating production rate on a plot of the
log of production vs cumulative recovery (Fig. 14). The first part of the curve
projects reserves by extrapolating the decline rate after the first fracturing treatment.
The second part of the curve projects reserves by extrapolating the decline rate after
the refracturing treatment.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 42 of 52

Fig. 14. Example of a production decline plot used to estimate recoverable reserves.

9.2.2 Estimation of the Present Fracture Geometry


The present fracture geometry can be estimated by making a log-log plot of pressure
buildup or reciprocal production versus time. The data can then be matched with
type curves. The matched curve can be used to project the future decline of the
actual production, provided the proper boundary has been chosen. Fig. 15 is an
example of a buildup data type-curve match. Fig. 16 is an example of a
performance data type-curve match.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 43 of 52

Fig. 15. Example of a buildup data type-curve match.

Fig. 16. Example of a performance data type-curve match.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 44 of 52

9.2.3 Determining the Average Reservoir Pressure


The average reservoir pressure for a depletion-drive reservoir can be determined
using the fracture length and conductivity (obtained from type-curve matching) and
the FORECAST module in the FracCADE software. The values of net pay, initial
reservoir pressure, porosity, saturation and well spacing must be known.

9.2.4 Estimating Production Response From Refracturing


Presently, the only method available to Dowell for estimation of production response
from a refracturing treatment is the FORECAST or FracNPV module in the
FracCADE software. The estimated average reservoir pressure and the new
fracture conditions must be used. An overestimate of the initial production will occur
because of the pressure gradients in the formation due to previous production.
However, longer term production (after 90 days) should be fairly accurate. A long
shut-in time prior to stimulation decreases the gradient and the negative effect on
production forecasting.

9.3 The Effects of Fractured Well Pressure Distributions on Design


A successful refracturing treatment will penetrate an area of greater reservoir
pressure.
Oil and gas production creates pressure gradients in the reservoir. Fig. 17 illustrates
possible pressure contours after production from a fractured well having an infinite-
acting conductivity fracture penetrating greater than 65% of the assigned drainage.
Assuming that a new fracture would have the same azimuth as the previous fracture,
this well would not be a good candidate for refracturing because the reservoir is
being drained as efficiently as possible (adequate penetration and conductivity).

Fig. 17. Pressure contours around high-conductivity propped fractures.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 45 of 52

Better reserve recovery may be obtained from infill drilling, locating the wells along
the boundary lines and fracturing the new wells as shown in Fig. 18. A longer
fracture in the new well will make the drainage more efficient. For this illustration,
the horizontal formation permeability is isotropic and the fracture azimuth is known.
Not knowing these conditions contributes to the economic failure in infill drilling and
refracturing (Fig. 19).

Fig. 18. The location of an infill well to improve recovery.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 46 of 52

Fig. 19. Possible problem is fracture azimuth is unknown.


The pressure contours for a well with a relatively short propped fracture (less than
200 ft) are shown in Fig. 20. A postfracture buildup-test on this well would indicate a
short fracture. A refracture treatment should be considered for this well. The
refracturing treatment should be designed to penetrate considerably further (a larger
treatment than the original) with infinite-acting fracture conductivity (probably
requiring higher conductivity).

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 47 of 52

Fig. 20. Pressure contours around a short fracture or a fracture with low conductivity.
It is possible that the initial propped penetration is greater than that indicated by
testing. Computer simulation of the initial fracture treatment using the FracCADE
software should be performed to determine if indications of a short effective fracture
may be caused by low proppant concentration (less than 1 lbm/ft2), low retained
permeability, or proppant settling near the fracture tip. If these are possible causes
for the short apparent penetration, the volume of fluid in the refracturing treatment
may be the same or less than the fluid volume in the original fracture treatment.
The short initial fracture may be a result of excessive fracture height-growth. Height
growth is indicated by a relatively flat net pressure on a Nolte-Smith plot. If a Nolte-
Smith plot is not available, the initial shut-in pressure observed at the end of the first
fracturing treatment can be compared to the initial shut-in pressure observed at the
beginning of the first fracturing treatment. If height growth is indicated, the refracture
treatment may not be economical, depending on the distance to barriers. A
diversion technique such as the INVERTAFRAC* or DIVERTAFRAC* service should
be considered prior to performing the refracturing treatment. Proppant from the
initial treatment can act as a diverting material.
Occasionally, well tests will indicate apparent fracture penetrations of only tens of
feet. These wells are prime candidates for refracturing because the short
penetration is probably due to polymer damage or proppant settling, resulting in
fracture conductivity values less than one.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 48 of 52

The examples shown in Fig. 17, Fig. 18, Fig. 19, and Fig. 20 described above are for
ideal isotropic formations. The pressure contours would be altered by the reservoir
anisotropies. Combining the anisotropies with structure, thickness, and permeability-
thickness maps of the reservoir will provide additional insight for evaluating
refracturing candidates.

9.4 Effects of Reservoir Pressure Changes


The effect of reservoir pressure on stress is shown in Fig. 21. Decreasing the
reservoir pressure will decrease the minimum horizontal stress. Fig. 21 is based on
the simple elastic theory using Poisson's ratio, the overburden stress and pore
pressure. Eq. 25 illustrates this relationship.

v(σ v − 2 p pore ) + p pore


σh =
1 − Fv (25)
Where:
σH = minimum horizontal stress (psig)
v = Poisson's ratio (dimensionless)
σv = overburden stress (psig)
ppore = pore pressure (psig)
F = ratio of the maximum to minimum horizontal stress (should not be
greater than 2).
Eq. 26 illustrates the change in stress due to change in pressure.

1 − 2v
∆σ h = ∆p (26)
1 − Fv
Where:
∆σh = stress change (psig)
∆p = pressure change (psig)
v = Poisson's ratio (dimensionless)
F = ratio of the maximum to minimum horizontal stress.
The pressure drop across the filter cake and reservoir used in the fluid-loss
calculation will also change (Eq. 27).
 1 − 2v 
∆p fc = ∆p − 1 (27)
 1 − Fv 

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 49 of 52

Where:
∆pfc = pressure drop across the filter cake and reservoir (psig)
∆p = pressure change (psig)
v = Poisson's ratio (dimensionless)
F = ratio of the maximum to minimum horizontal stress.
Based on typical values, a rule-of-thumb is that the decrease in minimum horizontal
stress will be approximately 75% of the decrease in reservoir pressure. The fluid-
loss pressure will increase by 25% of the absolute value of the pressure change.

Fig. 21. Horizontal stress due to Poisson's ratio and pore pressure.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 50 of 52

9.5 Refracture Treatment Design Considerations


The treatment design methodology for refracturing is essentially the same as the
treatment design methodology for fracturing. There are some additional
considerations. These are:
• fracture initiation
• fracture extension
• fracture containment
• screenout problems
• fracture-fluid recovery.

9.5.1 Fracture Initiation


An acid or water injection treatment is usually performed prior to a refracturing
treatment. Any fracture initiation problems should be apparent during this operation.
A primary concern is whether the new treatment will create a new fracture or initiate
in the old fracture. If a new fracture is not created, the proppant from the old fracture
will probably provide some diversion. A low-viscosity prepad may be desirable. The
low-viscosity fluid will penetrate the proppant pack and pressurize the fracture to
some magnitude above the minimum stress. This will cause the old fracture to
widen, in turn causing the proppant to move. The preferred movement is settlement
to the bottom as opposed to horizontal transport which may create a tip screenout.

9.5.2 Fracture Extension


The probability of moving a large volume of proppant is low. Fluid flow over the
proppant is easier than fluid flow through the proppant. The ratio of fluid flow-rate
above the proppant-pack to the fluid flow-rate through the proppant-pack can be
determined using Eq. 28.

4.5 × 10 9 w 3hof (28)


R=
Where: k f whep
R = fluid flow-rate ratio (dimensionless)
w = fracture width (in.)
hof = open height above the proppant-pack (ft)
kfw = fracture conductivity (md.ft)
hep = exposed height of the proppant-pack (ft).

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Design
Page 51 of 52

The fluid flow-rate above the proppant-pack in an open fracture with a width of 0.03
in., having one-tenth the height of an exposed propped fracture with 450 md-ft
conductivity will be 27 times greater than the fluid flow-rate through the proppant-
pack. A FGS simulation to estimate the net pressure for this rate with the height
being that of the unpropped fracture can be performed. However, accuracy may be
questionable because of different width constraints at the top and bottom of the new
fracture (Fig. 22).

Fig. 22. Fracturing above a previous proppant-pack.

DOWELL CONFIDENTIAL
Section 400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Design Dowell
Page 52 of 52

If the fracture height above the proppant-pack is small enough to cause a net
pressure large enough to force additional fluid through the proppant, the pressure
will be high enough to fracture out-of-zone above the fracture and should be higher
that that of the original treatment. The net pressure slope would also be flat.
Therefore, the current magnitude of stress is important.

9.5.3 Fracture Containment


If a well has produced for a considerable period of time, the reservoir pressure will
have decreased, reducing the minimum horizontal stress near the wellbore. The
stress contrast between the pay zone and barriers will have increased if the stress in
the barriers remains constant. This is beneficial in providing better containment of
the new fracture. If the old fracture experienced considerable downward height
growth and is partially filled with proppant, it may act as a diverter and the
refracturing treatment may be contained.

9.5.4 Screenout
Screenouts during refracturing in depleted zones have been attributed to increased
fluid loss because of increased pressure across the filter cake. The higher net
pressure caused by less fracture height can cause an even greater increase in
pressure across the filter cake.
Insufficient pad volume may also contribute to a screenout because of an estimated
fracture height error in the design. FracHite logs and stress logs based on initial
reservoir pressure should be modified to reflect the current conditions.
If pressure depletion decreases the fracture gradient in the pay zone, the result will
be less fracture height. If the original fracture height is used for design purposes, the
erroneous treatment design will indicate higher fluid efficiency; therefore a smaller
pad volume than actually required will be specified. The actual treatment will be less
efficient for the volume of fluid because the fracture will be contained within the
leakoff height.

9.5.5 Fracture-Fluid Recovery


Foamed or energized fluids will aid in fracture-fluid recovery. In one case, a
refracturing treatment may be performed after a well has been shut-in for weeks.
This is done to allow the reservoir pressure near the well to build-up and aid in the
fracture-fluid recovery. In another case, the well will be produced until the day of the
refracturing treatment to lower the reservoir pressure, increasing the stress contrast
for enhanced fracture containment.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 1 of 69

TREATMENT EXECUTION

1 Job Planning ............................................................................................................................ 4


1.1 Minimum Job Planning......................................................................................................... 4
1.2 Minimum Service Quality and Safety Standards.................................................................. 5
1.3 Operational Considerations ................................................................................................. 5
1.3.1 Mixing Fracturing Fluids ............................................................................................. 5
1.3.2 Equipment .................................................................................................................. 6

2 Safety ........................................................................................................................................ 7

3 Environment ............................................................................................................................. 7

4 Quality....................................................................................................................................... 7
4.1 Location Quality Assurance ................................................................................................. 7
4.2 Fracturing Fluid Kit............................................................................................................... 8
4.3 Testing References.............................................................................................................. 8

5 Mixing and Metering of Additive Solutions ........................................................................... 9


5.1 Liquid Additives .................................................................................................................... 9
5.1.1 Mixing Liquid-Additive Blends................................................................................... 11
5.2 Dry Additives in Solution .................................................................................................... 11
5.2.1 Soluble Additives ...................................................................................................... 11
5.2.1.1 Mixing Solutions of Soluble Dry Additives...................................................... 12
5.2.2 Insoluble Additives.................................................................................................... 12
5.2.2.1 Mixing and Metering Fluid-Loss-Additive Slurries.......................................... 13
5.3 Liquid-Additive Injection and Metering ............................................................................... 13
5.3.1 Correcting the Additive Injection Rate to Proppant Concentration ........................... 14
5.3.2 Example Calculation for Mixing and Metering Additive Solutions ............................. 14
5.4 Water Solubilities of Various Dry Additives as a Function of Temperature........................ 15
5.5 Methanol Solubilities of L10 as a Function of Temperature............................................... 17
5.6 Mixing and Metering Parameters for Various Solutions ..................................................... 17

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 2 of 69

6 Pressure Analysis During Fracturing Operations ...............................................................22


6.1 DSP Downhole Sensor Package (Real-Time Downhole Data Acquisition) ........................22
6.1.1 Fracturing Applications..............................................................................................23
6.1.2 System Components.................................................................................................23
6.1.2.1 Pressure and Temperature Gauges...............................................................24
6.1.2.2 Interface Module.............................................................................................25
6.2 Other Real-Time Data Acquisition Systems .......................................................................25
6.3 Downhole Memory Gauges ................................................................................................25
6.4 Additional Equations...........................................................................................................25

7 A Methodology For Improving Computed Bottomhole Pressures ....................................26


7.1 Recommendations..............................................................................................................28
7.2 Discussion ..........................................................................................................................28
7.2.1 Calibration Phase......................................................................................................29
7.2.1.1 Field Calibration .............................................................................................29
7.2.1.2 Office Calibration............................................................................................31
7.2.2 Application Phase .....................................................................................................32
7.3 Procedure A  Estimating the Closure Pressure and the Perforation/Near-Wellbore
Friction Pressure .....................................................................................33
7.4 Procedure B  Estimating the Fluid Friction Pressure and the Net Pressure ...................34
7.5 Procedure C  Estimating the Fluid Friction of Slurry Using Rate Changes .....................35
7.6 Procedure D  Estimating the Slurry Friction Pressure Using the PPR Software .............36
7.7 Real-Time Application of Procedure A, Procedure B and Procedure C .............................37
7.8 Field Example.....................................................................................................................37
7.8.1 Net Pressures From Uncalibrated Field Data ...........................................................37
7.8.1.1 Estimating Closure Pressure and Perforation Friction....................................40
7.8.1.2 Measuring Pipe Friction of Non-Proppant Laden Fluids.................................42
7.8.1.3 Measuring Pipe Friction of Slurries ................................................................47
7.8.1.4 Office Calibration  Job Playback with PPR Software ..................................50
7.8.1.5 PPR Hydrostatic Head Computation ..............................................................60

8 Flowback Recommendations ................................................................................................63


8.1 Energized Fluids Flowback Procedure ...............................................................................63

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 3 of 69

8.2 Choke-Size Determination ................................................................................................. 64


8.2.1 Choke-Size Determination for Foam Flowback (Greater than 55-Quality) ............... 64
8.2.2 Choke-Size Determination for Energized (Nonfoamed) Fluids................................. 66

9 Contingency Plans................................................................................................................. 69
9.1 Insufficient Pump Rate ....................................................................................................... 69
9.2 Proppant Delivery Failure .................................................................................................. 69
9.3 Equipment Malfunction ...................................................................................................... 69
9.4 Screenout........................................................................................................................... 69
FIGURES
Fig. 1. Compatability of common Dowell fracturing additives..................................................... 10
Fig. 2. The DSP system. ............................................................................................................ 24
Fig. 3. Schematic of treatment with systematic rate changes. ................................................... 30
Fig. 4. Friction multiplier as function of sand concentration. ...................................................... 31
Fig. 5. Treatment data for example well  uncalibrated. ......................................................... 38
Fig. 6. Net pressure plot for example well  uncalibrated........................................................ 39
Fig. 7. Procedure A  Treatment data. ..................................................................................... 41
Fig. 8. Friction pressure of fresh water, 5 1/2-in. casing. ........................................................... 42
Fig. 9. Procedure B  Treatment data, YF540HT..................................................................... 43
Fig. 10. Field calibrated fluid friction curve, YF540HT. .............................................................. 45
Fig. 11. Treatment data during rate change, WF110. ................................................................ 46
Fig. 12. Field calibrated fluid friction curve, WF110 . ................................................................. 47
Fig. 13. Treatment data during first shut-down with proppant.................................................... 48
Fig. 14. Treatment data during second shut-down with proppant.............................................. 49
Fig. 15. Treatment data from end of proppant to shut-down...................................................... 51
Fig. 16. Treatment data for example well  calibrated. ............................................................ 53
Fig. 17. Net pressure plot for example well  calibrated........................................................... 54
Fig. 18. Treatment data during rate change, WF110  calibrated............................................ 55
Fig. 19. Procedure B  Treatment data, YF540HT  calibrated.............................................. 56
Fig. 20. Treatment data during first shut-down with proppant  calibrated. ............................. 57
Fig. 21. Treatment data during second shut-down with proppant  calibrated......................... 58
Fig. 22. Treatment data from end of proppant to shut-down  calibrated. ............................... 59
Fig. 23. Comparison of calculated hydrostatic pressures. ......................................................... 61
Fig. 24. Treatment data for example well  improved hydrostatic pressure. ............................ 62
Fig. 25. Determining QZ............................................................................................................. 65
TABLES
Table 1. J218 ............................................................................................................................. 15
Table 2. J353 ............................................................................................................................. 15
Table 3. L10 ............................................................................................................................... 16
Table 4. M3 ................................................................................................................................ 16

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 4 of 69

Table 5. M117 ............................................................................................................................16


Table 6. L10 ...............................................................................................................................17
Table 7. Aqueous J218 Solutions...............................................................................................17
Table 8. L10/Methenol Solutions ................................................................................................18
Table 9. Aqueous L10 Solutions.................................................................................................18
Table 10. Aqueous J353 Solutions.............................................................................................19
Table 11. Aqueous M3 Solutions................................................................................................19
Table 12. Aqueous M117 Solutions............................................................................................20
Table 13. J84 and J418 Slurries.................................................................................................20
Table 14. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with M2 ...................21
Table 15. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with U28..................21
Table 16. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with J465 or J474
Respectively...............................................................................................................21
Table 17. Aqueous Breaker Aid J466 Solution ...........................................................................22
Table 18. Summary of Procedures A through D.........................................................................28
Table 19. Pipe Friction as Function of Rate YF540HT, 5-1/2 in. Casing ....................................44
Table 20. Comparison of Actual and Estimated Pipe Friction Proppant Laden Fluid .................50
Table 21. Choke Coeffiecient versus Choke Size ......................................................................66
Table 22. Required Flowrate to Gas-Lift Water ..........................................................................67
Table 23. Flow of Water through Chokes ...................................................................................68

1 Job Planning
A considerable amount of job planning is required to ensure a safe, successful and
profitable fracturing treatment. Historically, many on-location problems have been
directly attributed to an oversight in the planning phase.

1.1 Minimum Job Planning


Minimum job planning includes the following.
1. The wellsite should be visited prior to the job to determine the placement of
Dowell equipment and instrumentation, fracture tanks and contractor equipment.
Wellhead equipment should be surveyed to determine and verify treesaver and
treating adapter requirements.
2. A proppant-handling plan should be formulated.
3. A chemical plan should be formulated. This includes having adequate quantities
of chemicals (especially for reasonable last-minute changes), proper analytical
equipment and instrumentation.
4. An equipment plan should be formulated to ensure sufficient capacity and
adequate backup.
5. A pumping plan should be outlined, and planned pumping schedules provided to
appropriate personnel.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 5 of 69

6. Safety plans should be formulated. Safety-related job planning procedures are


provided in the Dowell Location Safety Standards.

1.2 Minimum Service Quality and Safety Standards


1. A prejob meeting will be held prior to departing the Dowell facilities.
2. The operating engineer/manager/supervisor is responsible for Dowell equipment
and personnel arrival at the jobsite in time to meet the client's time schedule.
3. A tailgate meeting will be held at the jobsite prior to job execution to ensure that
the client and Dowell personnel have a thorough understanding of job
procedures, safety procedures and individual responsibilities.
4. Treating equipment will be certified leak-free by a recorded pressure test prior to
the Dowell/client tailgate meeting.
5. Rigup will be in accordance with (applicable) Dowell Location Safety Standards
5, 5A, 5B, 9, 11, and 16.
6. Well pressure, pump rate and slurry density will be monitored and recorded
during execution.
7. Each job will be pumped in accordance with the FracCADE* design.
8. A neat, comprehensive well treatment report is required for each job.
9. Service quality will be discussed and reviewed with the client at the wellsite.
10. The wellsite will be clean prior to departure.

1.3 Operational Considerations

1.3.1 Mixing Fracturing Fluids


Fracturing fluids may be batch-mixed or continuously mixed. Batch mixing has
slightly different meanings, depending on the fluid being prepared.
Batch Mixing Oil-Base Fluids
All of the components and additives (except breakers and fluid-loss additives) of an
oil-base fluid are blended together in the fracture tanks before pumping begins.
Often, the fluid is mixed the day prior to pumping because the gel takes several
hours to form. A breaker and a fluid-loss additive are added continuously as the fluid
is pumped. These materials are not premixed because the breaker can reduce the
fluid viscosity prior to pumping and the fluid-loss additive may settle in the fracture
tank.
Batch Mixing Water-Base Fluids
Most of the components and additives (for example, potassium chloride, antifoam
agent, bactericide, clay control agent, surfactant and polymer) are mixed in the water
* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 6 of 69

prior to pumping. The polymer is given sufficient time to hydrate in the tanks before
the treatment begins. If the fluid will be crosslinked, the pH value of the fluid is
adjusted for optimum crosslinking. In the case of a YF * 300, YF400, YF500, or
YF600 series fluid, the crosslinker is continuously added as the fluid is pumped.
Because crosslinking in a YF100 or YF200 fluid occurs at a high pH value, the
crosslinker is batch-mixed with the other components. An activator is continuously
added to raise the pH value as the fluid is pumped. The crosslinker solution for
YF100HTD fluids contains the crosslinker, activator and delay agent and is
continuously added.
Batch Mixing versus Continuous Mixing
Batch mixing affords unique opportunities for quality control. The exact properties
and uniformity for the base fluid can be determined prior to pumping operations.
Unfortunately, batch mixing also results in wasted materials. There are always “tank
bottoms,” the fluid that cannot be sucked out of the fracture tanks. Typically, tank
bottoms represent at least 7% of the total volume of fluid in the tanks, so 7% of the
batch-mixed materials are wasted. Environmentally, this can create a disposal
problem. If the treatment is terminated early, the remaining fluid is wasted and will
require disposal. Also, if the treatment is postponed and the fluid degrades due to
bacterial action, the entire quantity of fluid will have to be discarded.
Continuously-mixed fluid is more desirable from an economic standpoint. All
materials are added as the fluid is pumped into the well, so there is no wasted fluid
and no unnecessary expense. This type of operation requires accurate metering of
all materials and makes quality control difficult. Real-time rheology measurements
provide the best quality assurance. Continuous-mix also requires that the desired
chemical reactions such as polymer hydration (water-base fluids) or gel formation
(oil-base) fluids should rapidly occur. The PCM* precision continuous mixer is a
pumping and blending system specially suited to continuous mixing of water-base
fluids.

1.3.2 Equipment
Refer to Section P-7 of A Practical Companion to Reservoir Stimulation for a
discussion of the organization and planning process of equipment placement and
rigup.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 7 of 69

2 Safety
The following safety standards are applicable to wellsite safety concerns typical in
day-to-day operations.
• Standard 5, Location Safety
• Standard 5A, Oilfield Location Safety  Offshore Operations
• Standard 5B, Marine Operations
• Standard 9, Pumping CO2
• Standard 11, Pumping Nitrogen
• Standard 16, Location Tool Operation.
These standards are provided in the Dowell Location Safety Standards handbook.
Other standards may be applicable, depending on wellsite conditions. The Dowell
Safety and Loss Prevention Manual provides all safety and loss prevention
standards.

3 Environment
A Dowell policy is to conduct business in a manner that assures optimum protection
for the environment. This policy may be complied with by
• careful compliance with relevant laws and regulations
• waste reduction
• efficient use of natural resources.

4 Quality
Testing of the fracturing fluid containing all components and additives prior to mixing
the fluid is the only way to ensure that the fluid will have the expected properties.
Mixing and testing procedures and specifications for the various fracturing fluids are
provided in the Fracturing Materials Manual Fluids. The Fracturing Fluids Quality
Control Manual provides quality-control procedures for testing field water, fracturing
fluids, proppant and hydrochloric acid.

4.1 Location Quality Assurance


Refer to Section P-8 in A Practical Companion to Reservoir Stimulation for additional
information on quality assurance for the following.
• fluids
• water quality
• base-fluid viscosity

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 8 of 69

• crosslinked fluids
• gelled oils
• foamed fluids
• proppants
• material balance.

4.2 Fracturing Fluid Kit


A fracturing fluid quality-control kit is available from Oilfield Chemical Manufacturing.
The kit contains the equipment and chemicals necessary to perform the following.
• water analysis
• fracturing fluid preparation
• crosslinking of Widefrac (YF*) fluids
• compatibility tests
• foam stability test
• proppant testing
• hydrochloric acid testing
• vapor pressure of oils.
The Fracturing Fluids Quality Control Manual, available with the test kit, provides a
detailed description of the test methods, test procedures, interpretation of results and
troubleshooting.

4.3 Testing References


Reference material includes the following.
Water Analysis
• American Petroleum Institute (API) RP 45, Analysis of Oil-Field Waters, Second
Edition, American Petroleum Institute (1981).
• Water Analysis Handbook, Hach Company (1989).
• Standard Methods for the Examination of Water and Wastewater, Sixteenth
Edition, American Public Health Association (1985).
Fracturing Fluids
• API RP 39, Standard Procedures for Evaluation of Hydraulic Fracturing Fluids,
Second Edition, American Petroleum Institute (1983).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 9 of 69

Proppants
• API RP 56 Recommended Practices for Testing Sand Used in Hydraulic
Fracturing Operations, First Edition, American Petroleum Institute (1983).
• API RP 60 Recommended Practices for Testing High-Strength Proppants Used in
Hydraulic Fracturing Operations, First Edition, American Petroleum Institute
(1989).
Laboratory Techniques
• Stimulation Laboratory Procedures Manual, Dowell (1984).
• Reservoir Stimulation, Second Edition, Schlumberger Educational Services
(1989)

5 Mixing and Metering of Additive Solutions


Fracturing operations often require additives to be added to the fracturing fluid
continuously, or “on the fly”. At times this may mean preparing liquid solutions of dry
materials, diluting liquid concentrates, and combining various materials into single
additive blends. The guidelines outlined in this section are intended to assist in the
preparation and metering of additive materials.

5.1 Liquid Additives


Liquid additives are generally intended to be individually added to fracturing fluids.
For treatments using several additives, limited tank storage space or additive pump
capacity can create a need to blend two additives in concentrated form for injection
through a single liquid-additive system. For example, if a treatment calls for
EZEFLO* F75N Surfactant at 2 gal/1000 gal and Surfactant F56 at 1 gal/1000 gal,
the two could be combined in a volume ratio of two parts F75N and one part F56 and
metered at a rate of 3 gal/1000 gal.
Caution must be exercised to ensure that blended additives are compatible and
adverse chemical reactions will not occur when combining liquid additives. Fig. 1
illustrates the compatibilities of a number of common fracturing fluid liquid additives.
In each case, additive combinations were evaluated at three different concentrations
(33%, 50%, and 66%) and at two temperatures (75°F [29°C] and 35°F [2°C]).
Additive combinations which are compatible over these ranges of conditions should
be compatible in conditions most likely to be experienced during normal fracturing
operations.
Fig. 1 indicates the compatibility of different additives if mixed in their concentrated
forms (as explained in the footnotes).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 10 of 69

Compatibilities may be different when diluted with common carrier fluids at additive
concentrations of from 1 to 10 gal/1000 gal carrier fluid. Specific tests should be
conducted when comingling in the diluted forms.

Fig. 1. Compatability of common Dowell fracturing additives.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 11 of 69

Additive combinations are indicated as being compatible (C), incompatible (I) or


partially compatible (P) in Fig. 1. For those combinations indicated as being partially
compatible, an explanation is included detailing the conditions for compatibility. In
some instances a particular additive combination may be physically compatible, for
example, form a homogeneous mixture, but may be chemically incompatible.
Combinations of acids and bases or oxidizers and reducing agents are examples of
this condition. Blends of these additives would lead to a degradation of one or both
of the components and possible reactive chemical hazards. In all cases these
combinations are indicated as being incompatible.

5.1.1 Mixing Liquid-Additive Blends


Mixing requirements for preparation of additive blends depend on the viscosities and
relative densities of the materials involved.
• Low-Viscosity and Similar Densities
Blends of low-viscosity materials with similar densities can be mixed by
combining the additives in a tank and circulating at a rate sufficient to ensure a
homogeneous blend.
• High-Viscosity or Large Density Difference
For very-viscous materials or where a large difference in material densities exists,
more intense agitation is required. In these cases, blends should be prepared in
stirred vessels. The mixing vessel should be loaded with the less viscous or less
dense liquid, and the more viscous or more dense liquid added slowly with good
agitation. Mixing should be continued until a homogeneous blend is obtained.

5.2 Dry Additives in Solution


Dry additives may be soluble or insoluble.

5.2.1 Soluble Additives


The most efficient method of adding soluble, dry additives to the fracturing fluid is to
prepare a liquid concentrated solution of the additive or additives and meter using a
liquid-additive system. Table 1 through Table 5 provide the maximum water
solubility versus temperature of most common dry additives used in fracturing
treatments. Consideration must be given not only to the mix-water temperature but
also to storage and use temperatures in determining the concentration to be used
when preparing these solutions.
Table 6 provides the solubility of Crosslinker L10 in methanol. Because of the low
solubility of L10 in water, it may be more convenient to use the more concentrated
methanol solutions. These solutions are unusual in that the L10 solubility increases
slightly at lower temperatures making them usable at concentrations up to 20% by wt
over the entire temperature range likely to be encountered.
Mixing and metering parameters have been compiled to assist in the preparation and
use of the additive solutions in Table 1 through Table 6 and are provided in Table 7

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 12 of 69

through Table 12. Table 13 through Table 17 provide mixing and metering
parameters for fluid-loss additive slurries, activator slurries and breaker-aid solutions.
Mixing ratios in terms of pounds of additive per volume of water (or diesel) are given,
followed by the solution characteristics for each additive (explained below).
• Concentration of Additive in Solution: This is expressed three ways  weight %,
pounds of additive per gallon of solution and pounds of additive per pound of
solution.
• Final Solution Volume: The volume after addition of the indicated weight of
additive to the specified volume of fresh water.
• Solution Specific Gravity (SG): The specific gravity of the solution.
To use the data in the tables, the maximum allowable additive concentration must
first be determined based on the anticipated temperature conditions. From this
concentration, the appropriate mixing ratio can be selected and scaled to provide
sufficient quantity of additive and volume of solution. An example calculation is
provided in Section 5.3.2 of this document.

5.2.1.1 Mixing Solutions of Soluble Dry Additives


Most dry-additive solutions can be prepared by circulation, provided reasonably high
circulation rates are used. At high additive concentrations, particularly in the case of
L10 solutions, agitated vessels should be used to ensure complete and rapid
dissolution of the additive. The addition of Breaker J218, and to a lesser extent,
Potassium Chloride M117 and High Temperature Gel Stabilizer J353, to water
results in an endothermic heat of solution which will lower the solution temperature.
Under some circumstances, for example, high additive concentration or cold mix
water or both, this may lead to difficulty in dissolving these materials. In these cases
lukewarm mix water should be used. J218 solutions should not be prepared in hot
water.

5.2.2 Insoluble Additives


The most common insoluble additives used in fracturing treatments are inert fluid-
loss additives such as J84 and J418. These materials are added dry through a dry-
additive kit or as a concentrated slurry in a Waterfrac (WF) fluid. Mixing and
metering parameters for J84 and J418 slurries are provided in Table 13.
The additive slurry can be prepared with either guar or hydroxypropylguar (HPG)
gelling agents, and is most conveniently done using HPG Polymer Slurry J876 or
PSG Polymer Slurry J877. The recommended slurried polymer concentration is
13.5 gal/1000 gal (equivalent to WF160 or WF260). Below this concentration, rapid
settling of the insoluble additive occurs and at higher concentrations the final slurry
viscosity becomes quite high. A compatible bactericide, generally the same used in
the treating fluid, should be included to ensure slurry stability.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 13 of 69

The additive slurry viscosity and stability are also dependent on the fluid-loss
additive concentration. The recommended concentration of fluid-loss additive is in
the range of 30 to 50% (wt/wt). Higher levels lead to high slurry viscosity while lower
levels result in particle settling. Slurries that are within the recommended
concentration range will exhibit initial (24-hr) settling rates of 2% or less. Fluid-loss-
additive slurries should be prepared on a job-by-job basis to avoid long-term storage.

5.2.2.1 Mixing and Metering Fluid-Loss-Additive Slurries


Mixing of fluid-loss-additive slurries requires intense agitation to ensure proper
suspension of the solids. The slurry can be mixed in tanks equipped with paddle-
type mixers using the following procedure.
1. Mix the WF fluid (see recommended concentration above) and allow to hydrate.
2. Add the fluid-loss additive slowly while operating the mixer at high speed.
3. Continue agitation for a minimum of 30 min to ensure a homogeneous mixture.
Obtaining proper suspension of the material is critical.
Metering of fluid-loss-additive slurries should only be done using metering pumps
specially equipped to provide an oil flush of the shaft bearings. Without this oil flush,
the solids can enter the bearings resulting in rapid damage. Maintaining moderate
agitation of the slurry while metering prevents settling and ensures a uniform
composition.

5.3 Liquid-Additive Injection and Metering


The concentration of conventional liquid additives is stated in gal/1000 gal of
fracturing fluid. For these materials, the concentration is also the Liquid-Additive
Metering Rate. Dry-additive concentrations are stated as lbm/1000 gal of fracturing
fluid. When these materials are dissolved and added in solution, the metering rate
for the solution must be calculated. Once the liquid-additive metering rate or
additive-solution metering rate is known, the injection rate for the liquid is dependent
on the pump rate of the fracturing fluid (without proppant) and must also be
calculated.
The Additive-Solution Metering Rate (gal/1000 gal) is determined by dividing the
required dry-additive concentration (lbm/1000 gal) by the concentration of additive in
solution (lbm/gal).
The Liquid-Additive Injection Rate or Additive-Solution Injection Rate (gal/min) is
determined using Eq. 1.
q a = 0.042ca q (1)
Where:
qa = injection rate of liquid-additive or additive-solution (gal/min)
ca = liquid-additive or additive-solution metering rate (gal/1000 gal of fracturing fluid)
q = pump rate (bbl/min of fracturing fluid).

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 14 of 69

5.3.1 Correcting the Additive Injection Rate to Proppant Concentration


Additive concentrations are stated in lbm/1000 gal or gal/1000 gal of fracturing fluid.
Some blending systems monitor the flow rate of fracturing fluid (rate into the blender)
as well as the flow rate of the fluid containing proppant (rate out of the blender). If
the flow rate of fluid with proppant is used as the basis for determining additive
injection rates, calculation of the additive injection rate must account for the volume
of proppant to be correct.
The injection rate for a liquid additive or additive solution in proppant-laden fluids is
determined using Eq. 2.
0.3503 ca q d SG p
qa = (2)
8.32 SG p + C p

Where:
qa = injection rate of liquid-additive or additive solution (gal/min)
ca = liquid-additive or additive-solution metering rate (gal/1000 gal of fracturing fluid)
qd = Fracturing fluid pump rate with proppant (bbl/min)
SGp = specific gravity of proppant
Cp = pounds proppant added (PPA) per gal of fracturing fluid.

5.3.2 Example Calculation for Mixing and Metering Additive Solutions


This example demonstrates the calculations and use of Mixing and Metering Tables
to determine metering rate (gal/1000 gal) and solution injection rate (gal/min) for
additive solutions.
Given
• 148,000 gal of fracturing fluid with 8 PPA of sand.
• Ambient temperature of 60°F (16°C).
• J353 added at 10 lbm/1000 gal.
• Pump rate of 20 bbl/min.
then
1. The maximum concentration of J353 that can be used at 60°F (16°C) is 50.2%
by wt (Table 2).
2. Any J353 solution up to 800 lbm/100 gal of fresh water can be used (Table 10).
To be on the safe side, a concentration of 750 lbm/100 gal is chosen.
3. The total weight of J353 required is 1480 lbm. Since J353 is supplied in 100-lbm
bags, it will be more convenient to use 15 bags or 1500 lbm. This will also
provide a few additional gallons of solution to compensate for possible metering
inaccuracies.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 15 of 69

4. The volume of water required for the J353 solution is 200 gal (weight of J353
divided by the weight of J353 per 100 gal of fresh water times 100). Thus the
J353 solution is prepared by dissolving 1500 lbm of J353 in 200 gal of fresh
water.
5. The final solution volume is 296.2 gal (each 100 gal of fresh water with 750 lbm
of J353 produces a solution volume of 148.1 gal (Table 10).
6. The solution metering rate is 2 gal/1000 gal of fracturing fluid (J353
concentration of 10 lbm/1000 gal divided by weight of J353 per gallon of solution
which is 5.06 lbm/gal).
7. The solution injection rate is 1.23 gal/min (Eq. 2).
(0.35)( 2)( 20)( 2.65)
qa = = 1.23 GPM
( 8.32)( 2.65) + ( 8 )

5.4 Water Solubilities of Various Dry Additives as a Function of Temperature

Table 1. J218
Temperature (°F [°C]) Maximum Solubility (wt %)
32 (0) 30.4
40 (4.4) 33.1
50 (10.0) 36.3
60 (15.5) 39.1
70 (21.1) 41.6
80 (26.6) 43.9
90 (32.2) 45.8
100 (37.8) 47.5

Table 2. J353
Temperature (°F [°C]) Maximum Solubility (wt%)
32 (0) 44.5
40 (4.4) 46.0
50 (10.0) 48.0
60 (15.5) 50.2
70 (21.1) 52.6
80 (26.6) 55.2
90 (32.2) 58.0
100 (37.8) 61.0

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 16 of 69

Table 3. L10
Temperature (°F [°C]) Maximum Solubility (wt%)
32 (0) 2.67
40 (4.4) 3.03
50 (10.0) 3.56
60 (15.5) 4.19
70 (21.1) 4.91
80 (26.6) 5.74
90 (32.2) 6.66
100 (37.8) 7.68

Table 4. M3
Temperature (°F [°C]) Maximum Solubility (wt%)
32 (0) 6.5
40 (4.4) 8.1
50 (10.0) 10.8
60 (15.5) 14.5
70 (21.1) 19.0
80 (26.6) 24.4
90 (32.2) 30.7
100 (37.8) 37.9

Table 5. M117
Temperature (°F [°C]) Maximum Solubility (wt%)
32 (0) 21.9
40 (4.4) 22.8
50 (10.0) 23.8
60 (15.5) 24.8
70 (21.1) 25.8
80 (26.6) 26.7
90 (32.2) 27.6
100 (37.8) 28.4

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 17 of 69

5.5 Methanol Solubilities of L10 as a Function of Temperature

Table 6. L10
Temperature (°F [°C]) Maximum Solubility (wt%)
32 (0) >20.0
77 (25) 22.8
Note: K46 has a specific gravity of 0.79.
Caution: Methanol is a flammable liquid. Vapors may ignite when an ignition source
is contacted. Keep away from heat, sparks and flame.
Methanol is incompatible with aluminum. Mix and store in containers
constructed of materials other than aluminum.

5.6 Mixing and Metering Parameters for Various Solutions

Table 7. Aqueous J218 Solutions


J218 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/25 gal) (gal)
11 5.0 0.43 0.050 25.5 1.032
22 9.5 0.84 0.095 26.1 1.054
33 13.7 1.23 0.137 26.7 1.075
44 17.4 1.59 0.174 27.4 1.096
55 20.9 1.94 0.209 28.1 1.116
66 24.0 2.28 0.240 28.3 1.135
77 27.0 2.60 0.270 29.5 1.154
88 29.7 2.90 0.297 30.2 1.172
99 32.2 3.19 0.322 31.0 1.189
110 34.5 3.47 0.345 31.7 1.205
121 36.7 3.74 0.367 32.3 1.221
132 38.8 4.00 0.388 33.0 1.236
Caution: J218 must be prepared, stored and metered using equipment constructed
of corrosion-resistant materials such as plastic, glass or stainless steel.
J218 is an oxidizer and must not be mixed with oil or alcohol.
J218 solution degrades and must be used within 48 hours. Exposure to
noncorrosion-resistant equipment accelerates degradation and reduces
shelf life.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 18 of 69

Table 8. L10/Methenol Solutions


L10 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/25 gal) (gal)
6 2.80 0.189 0.028 31.8 0.810
12 5.44 0.373 0.054 32.2 0.821
18 7.95 0.552 0.079 32.6 0.833
24 10.32 0.726 0.103 33.1 0.843
30 12.58 0.896 0.126 33.5 0.854
36 14.72 1.060 0.147 34.0 0.863
42 16.77 1.220 0.168 34.4 0.873
48 18.71 1.376 0.187 34.9 0.881
54 20.57 1.526 0.206 35.4 0.890
Caution: Methanol is an inflammable liquid. Vapors may ignite when an ignition
source is contacted. Keep away from heat, sparks, or flame.
Methanol is not compatible with aluminum. Mix and store in containers
constructed of materials other than aluminum.

Table 9. Aqueous L10 Solutions


L10 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
6 0.71 0.060 0.007 100.4 1.003
12 1.42 0.119 0.014 100.7 1.007
18 2.11 0.178 0.021 101.0 1.011
24 2.80 0.237 0.028 101.3 1.015
30 3.47 0.295 0.035 101.6 1.019
36 4.14 0.353 0.041 101.9 1.023
42 4.79 0.411 0.048 102.3 1.027

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 19 of 69

Table 10. Aqueous J353 Solutions


J353 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
50 5.7 0.49 0.057 101.7 1.042
100 10.7 0.95 0.107 105.0 1.066
150 15.2 1.39 0.152 108.3 1.089
200 19.3 1.79 0.193 111.6 1.111
250 23.1 2.18 0.231 114.9 1.131
300 26.5 2.54 0.265 118.2 1.150
350 29.6 2.88 0.296 121.5 1.168
400 32.4 3.21 0.324 124.8 1.186
450 35.1 3.51 0.350 128.1 1.202
500 37.5 3.80 0.375 131.4 1.217
550 39.7 4.08 0.397 134.7 1.232
600 41.8 4.35 0.418 138.1 1.245
650 43.8 4.60 0.438 141.4 1.258
700 45.6 4.84 0.456 144.7 1.271
750 47.4 5.06 0.473 148.1 1.282
800 49.0 5.28 0.490 151.4 1.294
850 50.5 5.49 0.505 154.8 1.304

Table 11. Aqueous M3 Solutions


M3 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
25 2.91 0.250 0.029 100.0 1.030
50 5.66 0.499 0.057 100.1 1.059
75 8.25 0.748 0.083 100.3 1.087
100 10.70 0.944 0.107 100.6 1.113
125 13.00 1.240 0.130 101.1 1.138
150 15.20 1.480 0.152 101.6 1.161

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 20 of 69

Table 12. Aqueous M117 Solutions


M117 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
25 2.91 0.247 0.029 101.2 1.018
50 5.66 0.489 0.057 102.3 1.036
75 8.25 0.725 0.083 103.5 1.053
100 10.70 0.955 0.107 104.7 1.069
125 13.00 1.180 0.130 105.9 1.085
150 15.20 1.400 0.152 107.2 1.101
175 17.30 1.610 0.173 108.4 1.116
200 19.30 1.182 0.193 109.7 1.130
225 21.30 2.030 0.212 111.0 1.144
250 23.10 2.230 0.231 112.3 1.157
275 24.80 2.420 0.248 113.6 1.170
300 26.50 2.610 0.265 115.0 1.183

Table 13. J84 and J418 Slurries


FLA Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
350 29.6 3.03 0.296 115.4 1.230
400 32.4 3.40 0.324 117.7 1.257
450 35.1 3.75 0.350 120.1 1.282
500 37.5 4.08 0.375 122.4 1.307
550 39.7 4.41 0.397 124.8 1.330
600 41.8 4.72 0.418 127.2 1.352
650 43.8 5.01 0.438 129.7 1.372
700 45.6 5.30 0.456 132.1 1.392
750 47.4 5.57 0.473 134.6 1.411
800 49.0 5.84 0.490 137.1 1.429
850 50.5 6.09 0.505 139.6 1.446

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 21 of 69

Table 14. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with M2
M2 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/100 gal) (gal)
16.7 2.0 0.17 0.02 100 1.02
Note: Breaker Aid J318 may be added to this solution.

Table 15. YF100, YF200 Crosslink Activator Aqueous Solution Prepared with U28
U28 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/95 gal) (gal)
5 2.0 0.17 0.02 100 1.02
Note: Breaker Aid J318 may be added to this solution.
Note: U28, 30% NaOH has a specific gravity of 1.33.

Table 16. YF100, YF200 Crosslink Activator Aqueous Solution


Prepared with J465 or J474 Respectively
J465 or J474 Concentration in Solution
Composition (wt %) (lbm/gal) (lbm/lbm) Final Slurry Solution SG
(gal No. 2 Volume (gal)
Diesel per 47
lbm J465 or
J474)
34 17.00 1.31 0.1685 35.9 0.93
47 12.75 0.96 0.1275 48.9 0.90
74 8.50 0.62 0.0852 75.9 0.87
155 4.25 0.30 0.0426 156.9 0.84
Note: The above data are based on diesel specific gravity (SG) = 0.82 (6.82
lb/gal). The weight of the slurry, weight percent of activator and slurry
specific gravity will vary with the specific gravity of the diesel. These
factors will have to be considered if the flowmeters used are based on
mass flow rate rather than volumetric flow rate.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 22 of 69

Table 17. Aqueous Breaker Aid J466 Solution


J466 Concentration in Solution
Makeup (wt %) (lbm/gal) (lbm/lbm) Final Volume Solution SG
(lbm/25 gal) (gal)
6.4 3.0 0.26 0.03 25.1 1.03

6 Pressure Analysis During Fracturing Operations


The evolution of net pressure during a fracture treatment is used to infer the shape
of the propagating fracture. Successful application of this technique is dependent
primarily on the accuracy of the fracture closure pressure and the bottomhole
pressure.
Fracture closure pressure and bottomhole pressures can be directly measured but
seldom are because of cost and operational inconvenience. Closure pressure is
usually approximated using “rules-of-thumb” or by guessing. Bottomhole pressures
are generally estimated by extrapolating surface pressure measurements.
A methodology for improving computed bottomhole pressures is provided in this
manual.
Treatment Evaluation provides a discussion of pressure interpretation during
fracturing operations.

6.1 DSP Downhole Sensor Package (Real-Time Downhole Data Acquisition)


The DSP* downhole sensor package is a tool system consisting of a group of
bottomhole sensors that provide real-time data acquisition. The sensor package is
run on coiled tubing and provides bottomhole pressure, bottomhole temperature and
casing collar locator data. The sensor inputs are processed and displayed on an
interface module screen. Data is stored in ASCII format for post-treatment analysis.
The DSP system has several advantages over downhole memory gauges.
• Data collection is in real-time, so the data can be used immediately.
• Since the sensors are connected to the surface, if a treatment or sensor anomaly
occurs, it is immediately known, not discovered after the treatment.
• Gauges, as well as data collection, are controlled from the surface so treatment
delays do not affect the recording of data, and there is no time limit to data
recording.
The DSP package does not have a telemetry system (a computer as part of the
bottomhole assembly allowing the exchange of digitized packets of processed data

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 23 of 69

both up and down the wireline). Data, in the form of electrical signals, is sent for
processing and recording from the sensors to the interface module on the surface.
The DSP system is designed to be used in conjunction with the PRISM* wellsite data
acquisition and recording system. Information is passed via wireline from the DSP
system temperature, pressure and casing collar locator tools to the interface module
and then to the portable acquisition computer. Depth data is acquired at the coiled
tubing unit, fed through to the portable acquisition computer, then on to the interface
module for use in processing bottomhole temperature, pressure and casing collar
data versus depth. Data is also sent to a laptop personal computer to create
real-time plots of DSP system data versus time.

6.1.1 Fracturing Applications


By running the DSP gauges on the end of coiled tubing as a “dead string,” the
bottomhole pressure can be used for real-time fracturing pressure analysis. The
bottomhole pressure can also be used to accurately evaluate the friction pressure of
the fracturing fluid in the coiled tubing annulus. The bottomhole temperature can be
used to adjust nitrogen rates, and therefore, bottomhole foam quality on foam
fracturing treatments. The bottomhole temperature can also be used to calculate
tubular stresses due to temperature changes.

6.1.2 System Components


The following discussion briefly describes the various components of the DSP
system. The Coiled Tubing Operators Manual contains a manual section for the
DSP system and provides a more detailed discussion. The DSP system is shown in
Fig. 2.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 24 of 69

Fig. 2. The DSP system.

6.1.2.1 Pressure and Temperature Gauges


There are two types of combination pressure and temperature gauges that can be
used with the DSP system.
1. The FTPT-CA (10,000 psi maximum) and the FTPT-DD (20,000 psi maximum).
These are available from Schlumberger (WTA). The operating temperature
range is from 32 to 350°F (0 to 177°C).
2. The EPG-520 (10,000 psi maximum), available from GRC.
GRC also offers a fast-response temperature tool, the ETG-512, which can be used
in conjunction with the EPG-520 pressure gauge. The FTPT gauges are
recommended and can be calibrated at any Schlumberger WTA district worldwide.
The GRC gauges can be calibrated only at a facility in the US and a facility in the
UK. The GRC gauges are not compatible with the casing collar locator instrument.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 25 of 69

6.1.2.2 Interface Module


The interface module contains the central processing unit, power supply, sensor
signal conditioning circuitry, analog-to-digital converter, communication interface,
display, and keypad. The interface module has the same outer appearance as the
standard remote display (SRD) in the portable acquisition computer (PAC) system;
however, the interface module is not interchangeable with the SRD.

6.2 Other Real-Time Data Acquisition Systems


Real-time bottomhole pressure data can be generated using a quartz gauge on a
side-port mandrel placed just above the packer. Typically, the wireline is banded to
the outside of the tubing (a slow process). Any gauge having ASCII or analog output
will work and quartz gauges that output analog voltage are available. In most cases,
ASCII output is all that is available because a quartz gauge frequency output
requires computer software to derive pressure. To read the ASCII data, a personal
computer equipped with two serial ports is required. One serial port reads the data
and the other serial port outputs the data (emulating a RDA) to the PPR*. The
software for this interface was developed by Guydon in Fort Worth, TX and is set up
for four different quartz system formats. If different gauges are used, software
changes will have to be made. The software is compiled, and although Dowell has
the source code, it calls a library utility that is not readily available; therefore, Guydon
has to make the changes.

6.3 Downhole Memory Gauges


Pressure gauges can be placed downhole (for example, in a gas-lift mandrel) to
record pressure during the fracturing treatment. After the fracturing treatment, the
gauge is retrieved using wireline, and the data is down-loaded and processed for
analysis.

6.4 Additional Equations


Perforation Friction Loss (Eq. 3)
p pf = pw − pisi − ptf (3)

Where:
ppf = perforation friction pressure (psi)
pw = surface fracturing pressure (psi)
pisi = initial shut-in pressure (psi)
ptf = tubular friction pressure (psi).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 26 of 69

Number of Open Perforations (Eq. 4)

qi2 p f
n= (4)
p pf d p f 4 ( 0.323)

Where:
n = number of open perforation
qi = constant pump rate (bbl/min)
ρf = density of the fracturing fluid (g/cm3)
ppf = perforation friction pressure (psi)
dpf = average diameter of perforation (in.).

7 A Methodology For Improving Computed Bottomhole Pressures


Various modes of fracture propagation can be inferred from straight-line periods
which develop on log-log plots of wellbore net pressure versus time. (See Treatment
Evaluation.) These modes characterize different fracture shapes.
Numerical simulators are used to augment the classical Nolte-Smith analysis. These
models are more representative than 2D fracture models because they account for
fracture height growth. The simulators are used to “history match” treatments by
systematically varying model parameters (for example, stress gradients and leak-off
coefficients) until predicted net pressures match actual net pressures. The geometry
of the simulated fracture is assumed to resemble the actual fracture if a reasonable
pressure match is obtained.
The successful application of any method of fracturing pressure analysis relies on
the accuracy of the fracturing pressure data. In particular, one must know the
bottomhole fracturing pressure and the magnitude of the least principal stress
(fracture closure pressure) before net pressure can be calculated. Errors in either of
these will result in incorrect net pressures and incorrect fracturing pressure analyses.
Fracture closure pressure can be measured using established methods; however,
for most treatments, closure pressure is extracted from “rules-of-thumb” or guessing.
Bottomhole pressure can be measured with bottomhole pressure gauges or “dead-
string” configurations. Unfortunately, these preferred techniques are rarely used
because of cost and operational inconvenience. Bottomhole pressures are usually
estimated by extrapolating surface pressure measurements. These extrapolations
are of little value unless perforation friction, fluid/slurry friction and hydrostatic
pressures are known. The degree to which the computed bottomhole pressures
resemble actual bottomhole pressures depends primarily on the accuracy of the
fluid/slurry friction data. The effect of fluid friction on computed bottomhole pressure
increases as the pipe diameter decreases; therefore, accurate fluid friction data are
especially necessary for treatments pumped down tubing.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 27 of 69

A methodology is provided in this document from which improved engineering


estimates of closure pressure, perforation friction, fluid/slurry friction, bottomhole
pressure and net pressure are obtained. This methodology uses standard Dowell
hardware and software and causes no inconvenience to the customer. A
comprehensive field example illustrating the methodology is also provided.
Net Pressure
Net pressure is a function of bottomhole pressure and fracture closure pressure. To
calculate accurate net pressures, bottomhole pressures and fracture closure
pressure must be known.
Fluid Friction Data
Fluid friction data provided in the Fracturing Materials Manual  Fluids can differ
significantly from field data. This is inevitable because of field variations in fluid
composition, polymer concentration and surface water temperature. These
variations affect fluid rheology and crosslink time. However, field calibrated fluid
friction data can be collected easily using the procedures provided in this document.
The use of these calibrated data significantly improves computed bottomhole and net
pressure estimates.
Estimating Perforation Friction
Procedure A provides a simple method for estimating perforation friction. This is
important because many apparent Dowell service failures can be traced to faulty
perforating practices. Procedure A provides a means to identify inadequate
perforations prior to a treatment.
Unplanned Shutdowns
Unplanned shutdowns can occur during any stage of a fracturing treatment. These
shutdowns are extremely valuable because they provide information for calibrating
fluid and slurry friction.
Proppant Friction Coefficient and Fluid Friction Gradient
The proppant friction coefficient (sfexp) and fluid friction (pf) gradients are consistent
among similar treatments for vertical wells. Therefore, once sfexp and pf are correctly
determined on one treatment, they can be used on subsequent treatments that use
the same fluid and are pumped down the same pipe size.
PPR Software
The PPR software routine for calculating hydrostatic pressure is deficient. When
proppant is ramped slowly and continuously over a long period of time, the computed
hydrostatic pressure has a “stair-stepped” character. Currently, the PPR software
only recalculates hydrostatic pressure when wellhead slurry density changes by 2%
or more.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 28 of 69

7.1 Recommendations
The procedures provided in this document can be used to generate field-calibrated
fluid friction data. These procedures consist of little more than brief injection rate
changes. They are easy to implement and cause no inconvenience to the client.
The calibrated friction data should be archived and used on subsequent treatments.
Procedure A should be used to estimate fracture closure pressure when more
rigorous techniques are not used. The procedure consists of injecting a small
volume of ungelled water into the formation and monitoring a brief pressure fall-off.
The procedure also provides a method for estimating perforation friction. This
procedure should be used prior to all treatments to identify problems relating to
perforations.
Treatment data recorded during unscheduled shut-downs should not be wasted.
These data should be used routinely to expand and enhance the district's fluid
friction database.

7.2 Discussion
The methodology consists of calibration and application phases. The calibration
phase is devoted primarily to measuring the pipe friction of the fracturing fluids (fluid
with and without proppant). This phase can be completed in as little as one fracture
treatment, provided fluid friction data are required for only one set of fluids in one
pipe size. During the application phase, the results of the calibration phase are
coupled with a brief on-site pump test to produce meaningful, real-time, net-pressure
plots. The components of each phase are summarized in following sections of this
document and their use demonstrated with an example.
Throughout this document, numerous references are made to Procedure A through
Procedure D. These procedures are the foundation of the methodology. The
objective of each procedure is summarized in Table 18.

Table 18. Summary of Procedures A through D


Procedure Objective
A Estimate fracture closure pressure (pc) and perforation friction (ppf)
just prior to the fracture treatment.
B Estimate pipe friction (ptf) of fluids without proppant using rate
changes during the fracture treatment.
C Estimate pipe friction (ptf) of fluids containing proppant using rate
changes during the fracture treatment.
D Estimate and/or confirm pipe friction (ptf) of fracturing fluids with and
without proppant using job playback with PPR software.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 29 of 69

7.2.1 Calibration Phase


If the bottomhole pressure is not measured directly during a treatment, the
bottomhole fracturing pressure must be estimated using Eq. 5. The PPR software
uses Eq. 5 to calculate the bottomhole fracturing pressure.
piw = pwh + ph − p pf − ptf (5)

Where:
piw = bottomhole fracturing pressure (psig)
pwh = surface treating pressure (psig)
ph = hydrostatic pressure of wellbore fluid/slurry (psig)
ppf = frictional pressure drop through perforations or near-wellbore restrictions (psig)
ptf = pipe friction pressure (psig).
For all Dowell fracturing treatments pwh is measured directly. ph is calculated by the
PPR software using input constants (for example, well depth and tubing/casing size)
and measured data (for example, injection rate and wellhead slurry density). ppf and
ptf are generally not measured, consequently, piw from Eq. 5 is usually questionable.
Eq. 5 will provide reliable estimates of piw if field calibrated values of ppf and ptf are
used.

7.2.1.1 Field Calibration


Perforation friction pressure, ppf and pipe friction pressure, ptf (ptf for fluids with and
without proppant) can be measured easily in the field. The field procedures for
measuring these parameters are thoroughly explained in Procedure A, B, and C.
(Procedure A also provides a technique for estimating fracture closure pressure.)
These procedures rely on Eq. 5 and consist of nothing more than brief injection rate
changes. If the rate changes are performed in a systematic manner, friction
pressures can be obtained directly from the resulting changes in surface pressure.
A schematic of a treatment that includes these systematic rate-changes is shown in
Fig. 3. The technique described in Procedure A corresponds to the time interval
labeled “A” in Fig. 3. Similarly, the techniques described in Procedure B and
Procedure C correspond to the intervals labeled “B” and “C” in Fig. 3. A summary of
the procedures is provided in Fig. 3.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 30 of 69

Summary of Procedures A through D


Procedure Objective
A Estimate fracture closure pressure (pc) and perforation friction (ppf)
just prior to the fracture treatment.
B Estimate pipe friction (ptf) of fluid without proppant using rate changes
during the fracture treatment.
C Estimate pipe friction (ptf) of slurries using rate changes during the
fracture treatment.
D Estimate and/or confirm pipe friction (ptf) of non-proppant and
proppant-laden fluids using job playback with PPR software.
Fig. 3. Schematic of treatment with systematic rate changes.
The successful execution of the field calibration phase provides the data necessary
to construct graphs of ptf versus injection rate (q) for a set of fluids in a given pipe
size (for an example, see Fig. 10). As mentioned earlier, the required data can be
obtained during one fracture treatment. However, because the procedures are quick

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 31 of 69

and easy, additional data should be collected on subsequent jobs to refine the “field
calibrated” fluid friction curves.

7.2.1.2 Office Calibration


The office calibration phase focuses on refining the pipe friction pressure (ptf)
estimates of the slurry. During this phase, the field calibrated friction data are used
in conjunction with the job playback features of the PPR software.
Tubular friction pressure of the slurry can be measured in the field using Procedure
C. However, the technique only yields ptf at a single proppant concentration or pump
time (for “stair-stepped” or “ramped” proppant schedules respectively). Therefore,
this procedure must be supplemented to account for the change in ptf with proppant
concentration.
The PPR software uses the Landel Correlation1 to model these changes. The
correlation, shown in Fig. 4, gives a friction multiplier which is the multiple increase in
friction of the slurry over that of the fluid without proppant.

Fig. 4. Friction multiplier as function of sand concentration.


ptf , slurry = fm ptf , base (6)

Where:
Ptf, slurry = frictional pressure drop of slurry through tubulars (psig)

1 Nolte, K.G.: “Fluid Flow Considerations in Hydraulic Fracturing,” paper 18537 presented at the 1988 SPE Eastern
Regional Meeting, Charleston, Nov. 1-4

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 32 of 69

fm = friction multiplier (dimensionless)


Ptf, base = frictional pressure drop of base fluid through tubulars (psig).
To use the correlation, the proppant friction coefficient (sfexp) must be estimated. The
default value for sfexp in the PPR software is 0.35.
There are two ways to obtain sfexp. The first is to use Procedure C. Procedure C uses
the pressure response from a rate change to explicitly calculate the friction multiplier.
Once the friction multiplier is known for a given proppant concentration, the
appropriate sfexp can be determined from Fig. 4. If Procedure C is omitted, sfexp can be
determined by trial-and-error with the PPR software. This technique is detailed in
Procedure D. It consists of playing back the treatment .RAW file with different
assumed values of sfexp. The correct value of sfexp produces relatively smooth and
continuous bottomhole fracturing pressure (piw) values during the transition from the
end of proppant to shut-down at the end of the treatment (for example, the period
from “D” to “E” in Fig. 3). To obtain the best possible data, it is strongly
recommended that Procedure C and Procedure D be used concurrently.
For vertical wells the sfexp and ptf gradients (that is, ptf /1000 ft versus injection rate) are
consistent among similar treatments. Therefore, once the sfexp and ptf gradients are
correctly determined on one treatment, they can be used on subsequent treatments
which use the same fluid and are pumped down the same pipe size.

7.2.2 Application Phase


The objective of the application phase is to produce meaningful, real-time, net
pressure plots. This is accomplished by combining the calibrated fluid friction data
with pre-job estimates of closure pressure and perforation friction pressure.
Procedure A details the techniques for estimating these parameters.
Closure pressure is not measured rigorously with Procedure A. However, the
estimated closure pressure value is superior to estimates obtained by guessing.
Also, it may be possible to develop a multiplicative “correction factor” relating the
closure pressure estimate from Procedure A to the actual closure pressure value.
(Eq. 7).

(
k = pc, proc A / Pc,actual ) (7)

Where:
k = correction factor relating pc,proc A and pc,actual (dimensionless)
pc,proc A = closure pressure estimate from Procedure A (psig)
pc,actual = actual value of closure pressure (psig).
The actual closure pressure can be measured by performing pump-in/flowback tests
during the field calibration phase. The resulting correction factor can be applied to
subsequent wells provided Procedure A was applied consistently and the formation
properties were similar among the wells.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 33 of 69

7.3 Procedure A  Estimating the Closure Pressure and the Perforation/Near-Wellbore


Friction Pressure
Procedure (see item “A” in Fig. 3)
1. Start with a fluid having a known friction
pressure (for example, ungelled water).
2. Inject 50 to 100 bbl into the formation. Inject at
a rate (q) equal to the planned fracturing
treatment pump rate or a rate not exceeding
the maximum surface pressure, whichever is
less. This will create a small fracture.
3. Just prior to shutdown, reduce the pump rate
to a very low rate (a single pump in low gear
for [depth(ft)/1000] sec). This should reduce
pressure surging. Use the pressure value at
the beginning of low rate for the initial shut-in
pressure (pisi).
Closure Pressure (pc) Calculation
Assuming an efficiency of approximately 10%,
the small fracture will close after a shut-in time
of approximately:

tc =
Pump time @ q
10
( min)
The surface pressure at time tc, is surface
closure pressure, psc. Bottomhole closure
pressure is the sum of hydrostatic pressure, ph
and psc.
pc = psc + ph
Hydrostatic pressure should be calculated to
the middle perforation depth. Use pc to
calculate the net pressure during the
treatment.
Perforation/Near-Wellbore Friction Pressure
(ppf) Calculation
Use the pressure difference (∆p) before and
after shut-in and the known pipe friction (ptf) of
the injected fluid to calculate ppf.
∆p = pmax − pisi
and,

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 34 of 69

p pf = ∆ p − ptf

7.4 Procedure B  Estimating the Fluid Friction Pressure and the Net Pressure
Procedure (see item “B” and “C” in Fig. 3)
1. After performing Procedure A, pump approximately
two tubing, casing or annular volumes of fracturing
fluid to cool the fluid column to near surface
temperature. If this volume is 3/4 of the pad fluid
volume or greater, pump 3/4 of the pad fluid volume
at fracturing rate.
2. Reduce rate q1 by some factor f. The new rate is q2.
q2 = f q 1
for example, q2 = 0.7 or 0.5q1
3. Determine the pressure difference ∆p1 caused by the
rate change.
4. As soon as a reliable pressure is obtained, shut
down and determine the initial shut-in pressure pisi
and ∆p2. (If pressure surging is expected, use the
shutdown procedure outlined in Procedure A)
5. As soon as a reliable pressures are obtained,
resume pumping at rate q1.
6. Determine ∆p3 from pressures before and after
injection. ∆p1 + ∆p2 should be about the same as
∆p3.
Equations For Calculating Unknowns
p f 1 ( at q1 ) = ∆ P1 + ∆ P 2 − p pf
p f 2 ( at q 2 ) = ∆ P 2 − f 2 p pf
pn = ~ 1.1 ( pisi + ph − pc )
Correction For Fluid Lost
If concerned about fluid leakoff during rate change
period (trc), add additional pad volume equal to:
(1-ef) q1 trc
Where:
trc = rate change time (min)
ef = efficiency from the FracCADE software (fraction)
q1 = injection rate (bbl/min)
Reset System  Maintain Records

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 35 of 69

Reset pipe friction data (ptf) in system to match points (q1, pf1 /L), and (q2, pf2 /L). Do not
include perforation friction (ppf) in plot of log net pressure versus log time since it will
probably be lost during proppant injection. Log-log plots of ptf /1000 ft versus rate should
be kept for each fluid type and pipe size.

7.5 Procedure C  Estimating the Fluid Friction of Slurry Using Rate Changes
(See Item “C” in Fig. 3)
Problem
To estimate BHP during proppant stages the pipe friction pressure of the slurry must
be known.

Making use of the nomenclature in the above figure, the following equations can be
derived:
(∆p1/∆pt) ≈ ( ∆p1,slurry/∆pt,slurry) (1-C)
or, rearranging:
∆ pt, slurry ≈ ∆ p1, slurry(∆ pt/∆ p1) (2-C)
Theory
During Procedure A and Procedure B (Items “A” and “B” in Fig. 3) the
perforation/near wellbore friction pressure and the pipe friction pressure of the pad
fluid was measured. If the slurry is the same as the pad fluid (or similar), a rate
change sequence similar to Procedure B can be used to estimate the pipe friction
pressure of the slurry. The figure and equations above illustrate the idea.
Procedure
1. Perform Procedure B prior to proppant injection.
2. During proppant injection, change rate to q2 as was done in Procedure B. Note
∆ p1, slurry.
3. Use Eq. 2-C to calculate ∆ pt, slurry, the total pipe friction for the slurry.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 36 of 69

4. Distribute several of these rate changes throughout the proppant stages. Use
these data in conjunction with Procedure D to develop field calibrated friction
curves for enhanced real-time net pressure analyses.

7.6 Procedure D  Estimating the Slurry Friction Pressure Using the PPR Software
(See Item “D” to “E” in Fig. 3)
Assumptions
1. BHP from “D” to “E” does not change significantly.
2. Friction-rate relationships are known for the slurry and flush fluid.
3. Perforation/near-wellbore friction pressure is known.
Problem
To estimate the BHP during the proppant slurry stages, the slurry hydrostatic
pressure, perforation/near-wellbore friction pressure and pipe friction pressure of the
slurry must be known. The hydrostatic pressure is calculated internally by the PPR
software. We assume that perforation/near-wellbore friction has been measured
prior to the job and does not change significantly. The objective then is to estimate
the pipe friction pressure of the slurry.
Theory
Given the above assumptions, the following equation can be written:
ptf,D = ps,D + ph,D - ppf,D - pw,E, (1-D)
Where:
ptf,D = pipe friction pressure of slurry at “D” (psig)
ps,D = surface treatment pressure at “D” (psig)
ph,D = hydrostatic pressure at “D” (psig)
ppf,D = perforation/near-wellbore friction pressure at “D” (psig)
pw,E = bottomhole pressure at “E”, equal to surface pressure at “E” + hydrostatic at
“E” (psig).
Define a friction multiplier which accounts for the increase in the pipe friction
pressure when proppant is added to a fracturing fluid:
fm = pf,slurry / pf,clean at a given rate q (2-D)
Where:
fm = friction multiplier (dimensionless)
pf,slurry = pipe friction pressure of the slurry (psi)
pf,clean = pipe friction pressure of fluid without proppant (psi).

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 37 of 69

The friction multiplier can be determined from Eq. 2-D, or from Fig. 4 if the proppant
friction coefficient (sfexp) is known.
Procedure
1. Load job data into OCS PPR software and create a .RAW file.
2. Make initial estimate for sfexp using Eq. 1-D, Eq. 2-D, and Fig. 4. Enter the value
into the PPR software. The proppant friction coefficient is stored in register 190
of “log data” parameters in the PPR software. The default value is 0.35.
3. Rerun the job using the PPR software and ensure that the BHP does not change
abruptly from “D” to “E”. If it does, choose a new value for sfexp and try again.
4. Calibrated sfexp values are used to enhance real-time net pressure plots on future
jobs.

7.7 Real-Time Application of Procedure A, Procedure B and Procedure C


1. Perform Procedure A as recommended. Enter the estimated closure pressure
(pc) into register 141 of PPR logdata parameters. Note the calculated value of
perforation friction pressure (ppf). If ppf is excessive (greater than 500 psi),
suggest reperforating. Do not allow the PPR to calculate ppf during the treatment
since it generally changes.
2. Perform Procedure B as recommended. After shutdown, quickly calculate ptf of
the injected fluid at rates q1 and q2. This calculation gives pf1 and pf2 (pipe friction
at rate q1 and rate q2). Divide pf1 and pf2 by depth to midpoint of perforations (pf1 /L,
pf2 /L). Enter the resulting pairs of data (q1 and pf1 /L), (q2 and pf2 /L) into the PPR
software using the option “Modify Fluid Friction Curves” under the “Calculation
Options Menu.” Resume pumping. With practice, the calculations and data input
can be completed in less than one minute.
3. Distribute rate changes throughout the proppant stages as described in
Procedure C. The pressure data from these rate changes are generally not used
in real-time; however, they are valuable datum points for the office calibration
phase (Procedure D).
4. If possible, flush proppant at the slurry rate. This will facilitate friction corrections
during the office calibration phase.

7.8 Field Example


In this section, the methodology will be illustrated using a field example.

7.8.1 Net Pressures From Uncalibrated Field Data


Treatment data for the example well are shown in Fig. 5 and Fig. 6. The upper
graph in Fig. 5 shows the measured surface pressures (STP) and calculated
bottomhole pressures (BHP) as functions of pump time. The lower graph shows the
proppant concentration and pump rate as functions of pump time. Fig. 6 shows the

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 38 of 69

net pressure plot that was displayed in the field while this job was being pumped.
The data for these plots came directly from the PPR .RAW file.

Fig. 5. Treatment data for example well  uncalibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 39 of 69

Fig. 6. Net pressure plot for example well  uncalibrated.


Water containing 2% (wt/wt) KCl, WF110, YF540HT, and YF535HT fluids were
pumped during this treatment. Water containing 2% (wt/wt) KCl was pumped from
2 to 7.5 min, WF110 was pumped from 9 to 36 min, YF540HT was pumped from
36 to 70 min, and YF535HT was pumped from 70 min until the end of the treatment.
These fluids were pumped down 10,110 ft of 51/2-in. casing at an average injection
rate of approximately 60 bbl/min. Approximately 1,700,000 lbm of 20/40-mesh
proppant (sand) were pumped during the treatment. This well was perforated from
9,850 ft to 10,110 ft with 202, 0.38 in. diameter perforations at 120° phasing.
The bottomhole pressures in Fig. 5 were generated by the PPR using Eq. 5. Tubular
friction data (ptf) were obtained from the Fracturing Materials Manual  Fluids. For
simplicity, the Field Engineer defined only two fluids in the PPR; WF110 and
YF540HT. The assumption was that the frictional properties of YF535HT and
YF540HT were similar (within 10 to 15%). Net pressures in Fig. 6 were calculated
using the bottomhole fracturing pressure (piw) from Fig. 5 and a closure pressure (pc)
of 7180 psig. The closure pressure was obtained assuming the fracture gradient
was 0.72 psig/ft, a common “rule-of-thumb” for the area. The perforation friction was
assumed to be zero since an abundance of perforations was present.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 40 of 69

This treatment is an excellent study case because all aspects of the calibration
phase were implemented: Procedures A through C were performed in the field and
Procedure D was performed in the office. Additionally, there were two complete
shutdowns while proppant was being pumped. These occurred from 108 to 125 min
and from 140 to 145 min. The shutdowns were unplanned but provided excellent
opportunities to measure the pipe friction pressure (ptf) of the slurry. (Treatment data
collected during unscheduled shutdowns should never be wasted. These data
should be used routinely to enhance the local fluid friction database.)
The data obtained from Procedures A through D are individually displayed and
analyzed in the following sections. However, before proceeding, it is important to
note that the data obtained from Procedure A and Procedure B (that is, pc, ppf and ptf of
slurries) can be input into the PPR software and used real-time. In other words, one
does not have to wait until the post-job playback to use the data as was done in this
example. (This is the current practice of the district that provided this example
treatment.)

7.8.1.1 Estimating Closure Pressure and Perforation Friction


Fig. 7 shows the data that were obtained from Procedure A. These data were
recorded while water containing 2% (wt/wt) KCl water was injected down the casing.
Injection began at 10 bbl/min and continued until a small fracture was created and
extended (4.5 to 7.5 min). Injection was then stopped and a brief period of pressure
fall-off was recorded (7.5 to 9 min). The slight drop in calculated bottomhole
pressure at 7.5 min was caused by a brief time lag in the PPR calculation routine.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 41 of 69

Fig. 7. Procedure A  Treatment data.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 42 of 69

Following the guideline of Procedure A, for an efficiency of around 10%, the (small)
fracture closure time will be approximately one-tenth of the pump time. For this
case, the injection time into the fracture was ~3 min. Fracture closure would be at
~0.3 min, or ~20 sec after shut-down. The surface pressure at this time was
~2530 psig. This led to a bottomhole closure estimate of: 2530 psig + (9979 ft x 0.43
psig/ft) = ~6820 psig. From this result it is clear that the “rule-of-thumb” estimate of
7180 psig was not too good.
Procedure A was also used to calculate perforation friction. The surface pressure
just before shut-down was ~2900 psig. Just after shut-down the surface pressure
was ~2600 psig. Therefore, total friction while injecting at 10 bbl/min was (2900 psig
- 2600 psig) = 300 psi. From Fig. 8 (from the Fracturing Materials Manual  Fluids),
the pipe friction pressure (ptf) of water at 10 bbl/min down 51/2-in. casing is
30 psig/1000 ft. For the example well, ptf = 30 psig/1000 ft x 9979 ft = ~300 psi.
Consequently, the assumption of zero perforation friction appears reasonable.

Fig. 8. Friction pressure of fresh water, 5 1/2-in. casing.

7.8.1.2 Measuring Pipe Friction of Non-Proppant Laden Fluids


Fig. 9 illustrates the application of Procedure B using YF540HT. Note the stair-
stepped rate changes from 60 bbl/min to 32 bbl/min to 0 bbl/min and the relatively
small changes in calculated bottomhole pressures in the upper figure. (The rapid
drops in calculated bottomhole pressure at 41 min and 42 min are caused by time
lags in the PPR calculation routine.) In this case, the friction data from the Fracturing

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 43 of 69

Materials Manual  Fluids were close. When sudden rate changes produce little
change in the calculated bottomhole pressure, friction has been calculated correctly.

Fig. 9. Procedure B  Treatment data, YF540HT.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 44 of 69

Even though the friction data were close, they could still be improved slightly. At
60 bbl/min the surface pressure just prior to the rate change was ~3920 psig (40.5
min). After reducing the rate to 32 bbl/min the surface pressure was ~3340 psig
(41.5 min). Finally, when the rate was reduced to 0 bbl/min, the surface pressure
was ~2910 psig (43 min). These pairs of pressures and rates were used to calculate
the pipe friction pressure (ptf) of YF540HT. The results of the calculations are
provided in Table 19. This same procedure was performed in other wells by the
Dowell district that provided this example. Their field calibrated ptf data are provided
in Fig. 10 along with the book values.

Table 19. Pipe Friction as Function of Rate YF540HT, 5-1/2 in. Casing
1 2
Rate, bbl/min Surface Pressure, Pipe Friction, psi Friction Gradient,
psig psi/1000 ft
60 3920 1010 101
32 3340 430 43
0 2910 0 0
1
The numbers in this column were obtained by subtracting the surface pressure at
zero rate from the surface pressures at the rate of interest. Recall there was
essentially no perforation friction.
2
Numbers in this column were obtained by dividing pipe friction by depth to midpoint
of perforations (9979 ft/1000 ft = 9.979 kft).

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 45 of 69

Fig. 10. Field calibrated fluid friction curve, YF540HT.


Procedure B was not fully implemented during the injection of WF110; only a single
rate change was performed. Fig. 11 shows an expanded view of the treatment data
from 10 to 50 min. From 14 to 36 min, WF110 completely filled the casing. The
input friction data were obviously in error since the brief rate change at 15 min
caused an abrupt increase in the computed bottomhole pressure. In addition, as
YF540HT was started (36 min), and the PPR software switched its friction
calculations from the WF110 friction curve to the YF540HT friction curve, the
computed bottomhole pressures quickly rose to a nearly correct level (40.5 min).
Procedure B was performed fully in subsequent wells to generate friction data for
WF110. The field calibrated friction curve is shown in Fig. 12 along with the book
values.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 46 of 69

Fig. 11. Treatment data during rate change, WF110.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 47 of 69

Fig. 12. Field calibrated fluid friction curve, WF110.

7.8.1.3 Measuring Pipe Friction of Slurries


Since there were two complete shutdowns during the proppant stages, it was
possible to directly measure the pipe friction of the slurry. For obvious reasons,
Procedure C does not require a complete shutdown. However, because these
shutdowns occurred, it is possible to compare the actual friction values with those
obtained using Procedure C.
Fig. 13 shows an expanded view of the treatment data during the first complete
shutdown (average proppant concentration in the casing was 4.5-lbm proppant
added). The pump rate was stepped from 60 to 33 bbl/min and then to 0 bbl/min. If
the shut-down had not been necessary, the rate would have increased from
33 bbl/min back to 60 bbl/min, in accordance with Procedure C. The bottomhole
pressures in Fig. 13 are in error because they abruptly decreased when the rate was
changed (again, the dips in bottomhole pressure at 108.5 min and 111.5 min are
caused by lags in the PPR calculations). The data in Fig. 13 have been analyzed in

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 48 of 69

two ways. First, Procedure B was used since the shut-down data were available.
Then Procedure C was used assuming the complete shutdown did not occur. The
results of these analyses are shown in Table 20. Note that the friction estimate from
Procedure C is within 6% of the actual value.

Fig. 13. Treatment data during first shut-down with proppant.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 49 of 69

Fig. 14 shows expanded view of the treatment data during the second complete
shutdown (average proppant concentration in the casing was 5.1-lbm proppant
added). Again, when the rate was suddenly changed, the calculated bottomhole
pressures abruptly dropped, indicating that friction had been underestimated.

Fig. 14. Treatment data during second shut-down with proppant.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 50 of 69

From these two shutdowns it is possible to calculate the friction multiplier (fm). Recall
that fm is the ratio of the pipe friction of the slurry to that of the fluid without proppant.
Therefore, from the first shutdown, fm = 1150 psig/1010 psig = 1.14. The pressure
drop of 1150 psig was taken from Table 20 while 1010 psig was taken from Table
19. From the second shut-down, fm = (3090 psig - 1880 psig)/1010 psig = 1.20 (3090
psig and 1880 psig are the pressures just before and just after shut-down in Fig. 14).
The proppant friction coefficient (sfexp) can be calculated by rearranging the equation
for fm shown in Fig. 4. Since there were two shut-downs, two sfexp calculations were
performed. From the first shut-down, sfexp = 0.36, which is very close to the PPR
default value of 0.35. From the second shutdown, sfexp = 0.44. For the moment there
is no reason to favor one value over the other; therefore, an average value of 0.4
seems reasonable. During the office calibration phase, the validity of a particular sfexp
can be evaluated over the entire range of proppant concentrations.
Table 20. Comparison of Actual and Estimated Pipe Friction Proppant Laden Fluid
Actual Pipe Friction Procedure B
Rate, BPM Surface Pressure, psig Pipe Friction, psi
60 3200 1150
33 2580 530
0 2050 0
Estimated Pipe Friction  Procedure C
1 2 3
Rate, BPM Pad Surface Slurry Surface Pipe Friction of
Pressure, psig Pressure, psig Slurry, psi
60 3920 3200 1080
33 3340 2580 
0 2910  
1
This column was taken directly from the second column in .
2
“Slurry” refers to fracturing fluid with proppant.
3
Applying of Procedure C:
∆p1 /∆pt = ∆p1,slurry/∆pt,slurry
∆pt,slurry = ∆p1,slurry * ( ∆pt,pad/∆p1,pad)
∆pt,slurry = (3200-2580) * [(3920-2910)/(3920-3340)]
∆pt,slurry = 1080 psi (which is within 6% of actual)
7.8.1.4 Office Calibration  Job Playback with PPR Software
The office calibration phase focuses on refining the pipe friction estimates of the
slurry. Two constraints guide this process. The first constraint dictates that
computed bottomhole pressures cannot change abruptly when the injection rate
changes. The second constraint requires that the computed bottomhole pressures
remain relatively smooth and continuous during the transition from the end of
proppant to shut-down (“D” to “E” in Fig. 3).

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 51 of 69

During this example analysis, the computed bottomhole pressures changed abruptly
when the injection rate changed, thus violating the first constraint. Fig. 15 shows an
expanded view of the treatment data during the transition from the end of proppant
to shutdown. It is also clear that the second constraint was violated (note data from
231.5 to 236.5 min).

Fig. 15. Treatment data from end of proppant to shut-down.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 52 of 69

At shutdown, just after flush is completed, the bottomhole pressure is known. It is


simply the sum of the surface pressure and hydrostatic pressure. Computed
bottomhole pressures prior to shutdown should continuously approach this same
pressure value. In the example treatment, the bottomhole pressure just after shut-
down was 8040 psig (~237 min in Fig. 15). However, by extrapolating the best fit
line through the data from 220 to 232 min, it is clear that the computed bottomhole
pressures were approaching 8250 psig. This discrepancy indicates that friction
pressures were ~210 psig too low.
Two things can be done to increase the computed friction: 1) increase the pipe
friction values of the base fluid, or 2) increase the value of the proppant friction
coefficient, sfexp (which increases the friction multiplier, in turn increasing the pipe
friction). Earlier analyses (Fig. 9 and Section 7.8.1.2) found that the pipe friction
pressure (ptf) was nearly correct for the base fluid; therefore, the sfexp must have been
too low. This conclusion is consistent with the results obtained in Section 7.8.1.3
(that is, 0.36< sfexp <0.44).
The PPR software was again used to playback the treatment data for this well. This
time the data obtained from Procedures A through C were used in place of “rules-of-
thumb” or default data. Specifically, the data used were: closure pressure (pc) =
6820 psig, Fig. 10 and Fig. 12 for YF540HT and WF110 friction data, and sfexp = 0.4.
The results of the second playback are provided in Fig. 16 through Fig. 22.
Compare these figures with those given previously (that is, compare Fig. 16 with Fig.
5, Fig. 17 with Fig. 6, Fig. 18 with Fig. 11, etc.). Clearly the calibrated input data
have led to improved bottomhole and net pressure estimates.
In general, the pipe size determines the extent to which calibrated fluid friction data
will enhance computed bottomhole pressures. The corrections derived from such
data become more pronounced as the pipe size decreases. However, even for the
example well with 51/2-in. casing, the corrections were quite large during the WF110
stage (~1500 psi).

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 53 of 69

Fig. 16. Treatment data for example well  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 54 of 69

Fig. 17. Net pressure plot for example well  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 55 of 69

Fig. 18. Treatment data during rate change, WF110  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 56 of 69

Fig. 19. Procedure B  Treatment data, YF540HT  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 57 of 69

Fig. 20. Treatment data during first shut-down with proppant  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 58 of 69

Fig. 21. Treatment data during second shut-down with proppant  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 59 of 69

Fig. 22. Treatment data from end of proppant to shut-down  calibrated.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 60 of 69

7.8.1.5 PPR Hydrostatic Head Computation


The computed bottomhole pressures in Fig. 16 are an improvement over those in
Fig. 5; however, they are still fairly jagged and have noticeable stair-steps (for
example, stair-stepping from 65 to 110 min, and 195 to 230 min). Much of the
jagged pattern was caused by “noisy” surface pressure data. It was found that the
stairstepping was caused by the computed hydrostatic pressures from the PPR
software.
The PPR software will recalculate hydrostatic pressure when the wellhead slurry
density has changed by 2% or more. The consequences of this procedure are
shown in Fig. 23. The upper graph shows the computed hydrostatic pressures for
the example treatment. Again, these data were taken directly from the PPR .RAW
file. The stair-steps present in this figure coincide precisely with those in Fig. 5 and
Fig. 16. The lower graph shows what the computed hydrostatic pressures would
look like if recalculations were performed more frequently. In this case, the
hydrostatic pressure was recalculated when there was a 0.5% or greater change in
wellhead slurry density, or every minute, whichever occurred first. These
calculations were performed with a hybrid version of PPR 2.1 software that is not
available in the field. Fig. 24 shows the resulting bottomhole pressures. Note the
absence of stair-stepping.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 61 of 69

Fig. 23. Comparison of calculated hydrostatic pressures.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 62 of 69

Fig. 24. Treatment data for example well  improved hydrostatic pressure.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 63 of 69

8 Flowback Recommendations

8.1 Energized Fluids Flowback Procedure


Nitrogen or carbon dioxide or both are the gases most commonly used in foamed
and energized fluids. During flowback following a treatment, they provide an efficient
source of concentrated energy to aid in rapid and more complete post-treatment
cleanup. There are potential hazards associated with the use of nitrogen and carbon
dioxide. As the fluid exits the flowline during flowback, the gaseous phase expands
rapidly. This rapid release of energy must be controlled to avoid a loss of flowback
efficiency and to ensure personnel safety.
The recommended procedure for flowback of energized fluids is provided below.
Refer to Dowell Location Safety Standard 9, “Bleed Back of Well Pressure through
Dowell Connections” for additional information and rigup diagrams.
1. The pressure rating of all parts of the flowback hookup should be greater than
the expected shut-in pressure.
2. Connections containing “vee” threads should not be used because failure is
likely occur when subjected to lateral strain and cyclic vibration. Connections
containing round threads may be used, but flanges and square threads are
preferred whenever possible. Integral-union connections are preferable to any
pipe-thread connections.
3. Elbows and sharp turns should be avoided where possible. If a turn must be
made, use a tee with a solid plug.
4. A double-plug-valve arrangement should be used on the flowback line. Valves
should be as close to the wellhead as practical and no more than shoulder high.
5. A choke nipple is installed in the flowback line to control the flow. An adjustable
choke should not be used.
6. Use only rigid steel lines. Hoses and swivel joints may come loose and cause
serious problems.
7. All parts of the flowback line are to be restrained with stakes or cables or both to
prevent movement. Each joint of pipe on the ground should be staked. Any
supports used to elevate the line should also be staked to the ground.
8. The flowback line should be as straight as possible. At each change of direction,
the line must be adequately secured to withstand the backward thrust created as
the gas exits the pipe into the atmosphere.
9. If perforation ball-sealers are used, the well should be flowed back momentarily
and then shut in to allow the ball sealers to fall into the rathole.
10. If ball sealers are expected to flow back to the surface, a ball catcher should be
used to prevent plugging the choke assembly. The ball catcher should be
installed between the tee on the wellhead and the first valve. An additional

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 64 of 69

bleedoff tee may be installed on the ball-catcher assembly to facilitate removal of


the ball sealers.
If a ball catcher is not available, an additional pressure bleed-off tee and valve
must be installed upstream of the plug valves. If a ball sealer seats in the one of
the valves or in the choke nipple, the wellhead valve can be closed, pressure
relieved from the line, and the ball sealer removed.
A potential hazard, often overlooked, is asphyxiation. Nitrogen and carbon dioxide
can collect in low areas, displacing breathable air. Personnel should avoid these
areas and remain upwind at all times. Only one person should be in the vicinity of
the well during flowback operations. The use of remote-operated valves will
increase the margin of safety.
The actual flowback characteristics of a well can never be predicted. Unexpected
gas pockets or slugs of propping agents can cause problems at the surface.

8.2 Choke-Size Determination


The following guidelines may be used to determine the choke size. The guidelines
are intended to provide a starting point from which the most appropriate method for a
particular flowback situation can be determined.

8.2.1 Choke-Size Determination for Foam Flowback (Greater than 55-Quality)


1. Calculate P/T.
Where:
P = Upstream flowing pressure (psig)
T = Upstream nitrogen temperature (°R).
Note: The flowing pressure can be obtained from the wellhead gauge.
°R = °F + 460.
2. Use Fig. 25 to determine QZ,
Where:
Q = Minimum gas rate for tubing liquid flow (Mscf/D), and
Z = Nitrogen deviation factor.
Draw a line from the minimum gas velocity through the P/T value, to the pivot, then
through the tubing size to QZ.
3. Divide QZ by Z to find Q.
“Z” factors are provided in the Nitrogen Manual.
This gives the minimum nitrogen flowrate required to keep fluid moving up the tubing
or casing during flowback.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 65 of 69

4. Determine C using Eq. 8.


Q GT
C= (8)
P
Where:
C = choke coefficient required
Q = minimum gas rate for tubing liquid flow (Mscf/D)
G = gas specific gravity
T = upstream nitrogen temperature (°R)
P = upstream flowing pressure (psi).
If the temperature and specific gravity are unknown use 80°F (27°C) for the
temperature and a value of 1 for the specific gravity.
5. Determine the choke size from Table 21 using the value determined in step 4. If
the choke coefficient does not match one of the values in Table 21, use the next
larger value.

Fig. 25. Determining QZ.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 66 of 69

Table 21. Choke Coeffiecient versus Choke Size


Choke Coefficient Choke Size (in.)
3.21 3/32
6.25 1/8
14.44 3/16
26.51 1/4
43.64 5/16
61.21 3/8
85.13 7/16
112.72 1/2
178.74 5/8
260.99 3/4
Example
P is 1000 psi
T is 520°R (60°F, 16°C)
Casing size is 4.000-in. (ID).
1. P/T = 1000/520 = 1.9
2. See example on Fig. 25. QZ = 3,400 Mscf/D.
3. Q/Z = 3,400/0.996 = 3,414 Mscf/D.
3, 414 1 ( 520)
4. C = = 77.85
1,000
5. A choke coefficient of 77.85 corresponds to a choke size of 7/16-in.

8.2.2 Choke-Size Determination for Energized (Nonfoamed) Fluids


1. Based on pipe size, determine the minimum required flow rate (Table 22). This
is the minimum rate at which water will normally continue to flow up the tubing or
casing. Below this rate, the fluid will fall to the bottom of the well and will start to
load up the well.
2. Using Table 23, match the value for upstream pressure (from the wellhead
gauge) with the required flow rate in gal/min (determined in step 1). Read the
choke size from the top of the column. If the match is not close, use the next
larger choke size.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 67 of 69

Note: As the upstream pressure drops, the choke size will have to be increased to
maintain flow rate.
Example
Casing is 4.500-in.
Upstream pressure is 500 psi.
1. The minimum required flow rate is 28 gal/min (from Table 22).
2. From Table 23, the flow rate through a 1/4-in. choke at 500 psi is 25.4 gal/min.
The flow rate through a 3/8-in. choke is 57.2 gal/min. The correct choke size is
5/16-in. (larger than 1/4-in. but smaller than 3/8-in.)
Table 22. Required Flowrate to Gas-Lift Water

Pipe Size  OD (in.) Required Flowrate (gal/min)


2.000 10.5
2.375 14
2.875 17.5
3.500 21
4.500 28
5.500 35
7.000 45.5
2.375 in 4.500 12.25
2.375 in 5.500 17.5
2.875 in 5.500 14
2.875 in 7.000 28

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Execution Dowell
Page 68 of 69

Table 23. Flow of Water through Chokes


FLOW THROUGH CHOKE IN U.S. GALLONS PER MINUTE
Diameter of Choke (in.)
Upstream Velocity 1/64 1/32 1/16 1/8 3/16 1/4 3/8 1/2 5/8 3/4 7/8 1
(psi) through
Choke
(ft/sec)
1 7.44 0.0044 0.0178 0.0711 0.285 0.641 1.14 2.56 4.56 7.12 10.3 14.0 18.22
10 23.53 0.0140 0.0564 0.224 0.900 2.03 3.60 8.11 14.4 22.5 32.4 44.1 57.6
20 33.27 0.0198 0.0798 0.318 1.27 2.87 5.10 11.5 20.4 31.8 45.9 62.4 81.6
30 40.75 0.0242 0.0978 0.390 1.56 3.51 6.24 14.0 24.9 39.0 56.2 76.5 99.8
40 47.06 0.0279 0.113 0.450 1.80 4.05 7.20 16.2 28.8 45.0 64.9 88.3 115.
50 52.61 0.0312 0.126 0.503 2.02 4.54 8.06 18.1 32.2 50.4 72.5 98.7 129.
60 57.63 0.0343 0.138 0.550 2.20 4.95 8.83 19.8 35.3 55.1 79.4 108. 141.
70 62.25 0.0370 0.149 0.594 2.38 5.36 9.54 21.4 38.2 59.6 85.8 117. 153.
80 66.55 0.0395 0.159 0.635 2.55 5.73 10.2 22.9 40.7 63.6 91.7 125. 163.
90 70.58 0.0419 0.169 0.674 2.70 6.08 10.8 24.3 43.3 67.5 97.3 132. 173.
100 74.40 0.0442 0.178 0.712 2.85 6.41 11.4 25.6 45.6 71.2 103. 140. 182.
125 83.18 0.0494 0.199 0.794 3.19 7.16 12.7 28.6 50.9 79.6 115. 156. 204.
150 91.13 0.0542 0.218 0.871 3.49 7.86 14.0 31.4 55.9 87.3 126. 171. 223.
175 98.42 0.0586 0.236 0.940 3.77 8.50 15.1 33.9 60.2 94.3 136. 185. 241.
200 105.20 0.0626 0.252 1.01 4.04 9.09 16.1 36.2 64.5 101. 145. 197. 258.
225 111.6 0.0664 0.268 1.07 4.28 9.62 17.1 38.4 68.4 107. 154. 210. 274.
250 117.6 0.0700 0.282 1.13 4.50 10.1 18.0 40.5 72.1 112. 162. 221. 288.
300 128.9 0.0767 0.309 1.23 4.94 11.1 19.7 44.4 79.0 123. 178. 242. 316.
350 139.2 0.0828 0.334 1.33 5.32 12.0 21.3 48.0 85.3 133. 192. 261. 341.
400 148.8 0.0886 0.357 1.42 5.70 12.8 22.8 51.2 91.2 142. 205. 279. 365.
450 157.8 0.0940 0.379 1.51 6.05 13.6 24.2 54.4 96.8 151. 216. 296. 387.
500 166.4 0.0990 0.399 1.59 6.37 14.3 25.4 57.2 102. 159. 229. 312. 407.
600 182.2 0.1080 0.437 1.74 6.99 15.7 27.9 62.8 112. 174. 251. 342. 446.
700 196.9 0.1170 0.472 1.88 7.54 17.0 30.1 67.8 121. 188. 271. 369. 482.
800 210.4 0.1250 0.505 2.01 8.06 18.1 32.2 72.4 129. 201. 289. 394. 515.
900 223.2 0.1330 0.535 2.13 8.55 19.2 34.2 76.9 137. 213. 308. 419. 547.
1,000 235.3 0.1400 0.564 2.24 9.00 20.3 36.0 81.1 144. 225. 324. 441. 576.
1,200 257.7 0.1530 0.618 2.46 9.88 22.2 39.4 88.9 158. 247. 355. 484. 631.
1,400 278.4 0.1660 0.667 2.66 10.7 24.0 42.6 95.9 171. 266. 384. 522. 682.
1,600 297.6 0.1770 0.714 2.84 11.4 25.5 45.6 103. 183. 285. 410. 559. 730.
1,800 315.7 0.1880 0.756 3.01 12.1 27.2 48.3 109. 194. 302. 435. 591. 774.
2,000 332.7 0.1980 0.798 3.18 12.7 28.7 51.0 115. 204. 318. 459. 624. 816.
2,100 341.0 0.2030 0.817 3.25 13.1 29.4 52.2 117. 209. 326. 470. 639. 836.
2,200 349.0 0.2080 0.837 3.34 13.4 30.1 53.4 120. 214. 334. 481. 655. 855.
2,300 356.8 0.2120 0.856 3.42 13.7 30.8 54.6 123. 219. 341. 491. 670. 875.
2,400 364.5 0.2170 0.875 3.49 14.0 31.4 55.8 126. 223. 348. 502. 684. 894.
2,500 372.0 0.2210 0.892 3.55 14.2 32.1 57.0 128. 228. 356. 513. 698. 911.
3,000 407.5 0.2420 0.978 3.90 15.6 35.1 62.4 140. 249. 390. 562. 765. 998.
3,500 440.2 0.2620 1.050 4.20 16.9 38.0 67.4 152. 270. 422. 607. 823. 1,078.
4,000 470.6 0.2800 1.130 4.50 18.0 40.5 72.0 162. 288. 450. 649. 883. 1,152.
4,500 499.1 0.2970 1.200 4.76 19.1 43.0 76.5 172. 306. 478. 687. 935. 1,222.
5,000 526.1 0.3130 1.260 5.03 20.2 45.4 80.6 181. 322. 504. 725. 987. 1,290.
6,000 576.3 0.3420 1.380 5.50 22.0 49.5 88.3 198. 353. 551. 794. 1,078. 1,410.
7,000 622.5 0.3700 1.490 5.94 23.8 53.6 95.4 214. 382. 596. 858. 1,168. 1,528.
8,000 665.5 0.3950 1.590 6.35 25.5 57.3 102. 229. 407. 636. 917. 1,247. 1,629.
9,000 705.8 0.4190 1.690 6.74 27.0 60.8 108. 243. 433. 675. 973. 1,321. 1,730.
10,000 744.0 0.4420 1.780 7.12 28.5 64.1 114. 256. 456. 712. 1,025. 1,395. 1,822.

DOWELL CONFIDENTIAL
Section 500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Execution
Page 69 of 69

9 Contingency Plans

9.1 Insufficient Pump Rate


If the designed pump rate cannot be obtained after pumping one pipe volume, cease
pumping operations, recalculate basic fracturing pressures and discuss findings and
solutions with the client.
If diverting materials (for example, perforation ball sealers) were pumped prior to the
fracturing treatment (that is, as part of the treatment), the diverting materials may
have reseated in the perforations while pumping the pad fluid. The well should again
be flowed back momentarily and then shut-in to allow the diverting materials to fall
into the rathole.

9.2 Proppant Delivery Failure


The following options should be discussed with the client before the fracturing
treatment is started.
Stop pumping if the proppant cannot be delivered as designed and planned. The
only exception is if a screenout appears imminent in the last 20% of the job, then
consider flushing the pipe.
Proppant Delivery Failure During the First 1/3 of the Treatment
If the problem is rectifiable within thirty minutes, repair and continue as designed and
planned. If the problem is not rectifiable within thirty minutes, discuss with the client.
Proppant Delivery Failure After the First 1/3 of the Treatment
If the problem is rectifiable within thirty minutes, repair and continue as designed and
planned, increasing the pump rate 10% for a time increment equal to the shutdown
period. If the problem is not repairable within thirty minutes, discuss with the client.

9.3 Equipment Malfunction


Activate redundant systems as soon as possible and continue pumping as designed
and planned.

9.4 Screenout
Cease pumping operations when maximum allowable pressure is attained and
discuss the situation with the client. Obtain fluid and material samples.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 1 of 23

TREATMENT EVALUATION

1 Introductory Summary............................................................................................................. 2
1.1 Treatment Evaluation Methodology ..................................................................................... 2
1.2 Minimum Service Quality and Safety Standards.................................................................. 3

2 Fracturing Pressure Analysis ................................................................................................. 3


2.1 Injection Pressure Interpretation .......................................................................................... 3
2.1.1 Nolte-Smith Plot........................................................................................................... 3
2.1.2 In-Situ Stress Requirements........................................................................................ 7
2.1.3 References .................................................................................................................. 8
2.2 Pressure Decline Analysis ................................................................................................... 8
2.2.1 References .................................................................................................................. 9
2.3 Fracture Height Prediction and Post-Treatment Measurements.......................................... 9
2.3.1 Sonic Logs................................................................................................................... 9
2.3.2 References .................................................................................................................. 9

3 Treatment Performance Monitoring ..................................................................................... 10


3.1 Inverse Analysis of Treatment and Production Data Records ........................................... 10
3.1.1 Fracture Characterization Using the ZODIAC Software ............................................ 11
3.1.1.1 Fracture Storage............................................................................................ 12
3.1.1.2 Fracture Face Skin Damage .......................................................................... 13
3.1.1.3 Variable Fracture Conductivity....................................................................... 13
3.1.1.4 Reservoir Permeability Anisotropy................................................................. 14
3.1.1.5 Finite Reservoirs ............................................................................................ 14

4 Production Evaluation........................................................................................................... 22
4.1 References......................................................................................................................... 23
FIGURES
Fig. 1. Slope interpretation for the Nolte-Smith plot. .................................................................... 7
Fig. 2. Example of required stress contrasts................................................................................ 8
Fig. 3. High-conductivity fracture comparison. ............................................................................ 15

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 2 of 23

Fig. 4. Low-conductivity fracture comparison. ............................................................................15


Fig. 5. Fracture face skin damage, moderate-conductivity fracture............................................16
Fig. 6. Fracture face skin damage, finite-conductivity fracture. ..................................................16
Fig. 7. Fracture face skin damage comparison, low-conducitivy fracture...................................17
Fig. 8. Fracture skin damage comparison, low-conductivity fracture..........................................17
Fig. 9. Finite-conductivity fracturecomparison, high average dimensionless conductivity..........18
Fig. 10. Finite-conductivity comparison, moderate average fracture conductivity. .....................18
Fig. 11. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft. .............19
Fig. 12. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft. .............19
Fig. 13. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft. ....................20
Fig. 14. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft. ....................20
Fig. 15. ZODIAC software examples. .........................................................................................21
Fig. 16. ZODIAC software examples. .........................................................................................21
Fig. 17. ZODIAC software examples. .........................................................................................22

1 Introductory Summary
Treatment evaluation provides both the client and Dowell with a measure of how well
the design and execution were accomplished. The results also serve as an
experience data bank and reference point for future wells. Careful analysis and
appraisal of treatment data and production data will reduce well costs and improve
production by increasing efficiency, improving service, and applying the latest
technology.

1.1 Treatment Evaluation Methodology


Following the treatment execution, evaluation methodology is as follows.
1. Dowell treatment execution evaluation by the client and Dowell personnel:
• Dowell planning, mobilization, and rigup at the well site
• Performance and quality of Dowell materials
• Proportioning, mixing, and pumping of Dowell materials
• Recording and presentation of Dowell treatment execution data.
2. Evaluate the formation response relative to treatment design (for example,
design volume pumped or not, post-fracture logs for height similar to design
height).
3. Evaluate production response relative to the fracturing treatment (that is, actual
production response and predicted within acceptable limits).
4. Evaluate design methodology.

1.2 Minimum Service Quality and Safety Standards


1. Appointed Dowell personnel will hold monthly service quality reviews with clients

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 3 of 23

to discuss job execution and results.


2. A DESIGN-EXECUTION-EVALUATION* package will be presented to the client
within seven days of completion of the well.

2 Fracturing Pressure Analysis


Fracturing pressure analysis may be reduced to two distinct types of analysis.
1. Pumping pressure or injection pressure interpretation.
2. Pressure decline after the fracturing treatment.

2.1 Injection Pressure Interpretation


The basis for interpreting fracturing pressure is to use the net pressure (fluid
pressure inside the fracture). The behavior of net pressure with time (increasing,
constant, or decreasing) is related to the hydraulic fracture geometry (for example,
confined height, unconfined height, and rate of length/growth with time). The
magnitude of the net pressure is controlled by fracture geometry and the elastic
modulus of the formation and, to a lesser extent, by fluid viscosity and pump rate.
The most common analysis procedure is based on pressure behavior.

2.1.1 Nolte-Smith Plot


The Nolte-Smith plot (log-log plot) is used as a diagnostic tool to determine what is
actually happening in the fracture. By plotting bottomhole net pressure versus time,
the slope will indicate the mode of fracture behavior (Fig. 1).
Type I Behavior
Refer to Fig. 1. Increasing net pressure with a 1/8 to 1/4 slope indicates the fracture
is propagating with restricted height, unrestricted extension, that linear flow
dominates fluid loss, and that injection rate and fluid viscosity are constant. This is
type I behavior.
The effect of pressure is best seen by examining the continuity equation. That is,
Fluid Rate In (qin) = Loss Rate (qloss) + Rate of Fluid Storage in Fracture Volume
(∆w + ∆h + ∆L). Since pressure is proportional to fracture width, this can be rewritten
as qin = qloss + (a constant) (∆p + ∆h + ∆L).
This simple relation can provide valuable insight into what is occurring in the fracture.
In type I fracture behavior, qin is constant, so the right-hand side of the equation must
also be constant. Leakoff rate (qloss) is increasing with time as fracture area
increases. Net pressure and fracture length are increasing with time (so ∆p and ∆L
are > 0). Fracture height is constant (so ∆h = 0).
Now the effect of height growth can be considered. If fracture height starts to
increase, ∆h becomes positive and one, two, or all of the variables qloss, ∆p and ∆L must

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 4 of 23

decrease. There is no reason for new fracture height to decrease leakoff (in fact, the
opposite is more likely to occur) so ∆p or ∆L or both must decrease. If enough height
growth occurs, ∆p may go to zero and injection pressure will be constant. This is
type II behavior.
Type II Behavior
Refer to Fig. 1. Type II fracture behavior results in a flat net pressure response
indicating stable height-growth or that increased fluid loss is negating the predicted
pressure increases. Stable height-growth can be thought of as uniformly increasing
height (that is, height increasing at so many ft/min) and this height growth
compensates for the predicted pressure increase by gradually increasing the cross
sectional area of the fracture. This reduces the normal pressure increase by
reducing flow velocity in the fracture, and thus reducing the friction pressure drop in
the fracture. As height growth continues, the treatment may enter type IV behavior
(discussed below).
As discussed in type I behavior, the effect of height-growth can be seen from the
continuity equation. For a constant injection rate, an increase in fracture height must
be compensated for by a reduction in the rate of length extension or a reduction in
the rate of width (pressure) increase or both. This is also true for an increase in
fluid-loss rate.
The opening of small natural fissures intersected by the main fracture is one
mechanism for higher fluid loss. Normally, these fissures have higher permeability
than the matrix and the fluid can readily penetrate into the fissures and maintain a
pressure nearly equal to the pressure in the primary fracture. When the total
pressure in the fracture exceeds the stress holding the fissures closed, they open,
thereby increasing fracture volume and fluid-loss rate. This decreases the pressure
in the fracture and allows the fissures to close. Pressure in the main fracture then
increases and the cycle is repeated. The fissures act as a “pressure regulator”
maintaining a constant injection pressure. A significant portion of the injected fluid
can be lost in this process because of the large number of fissures that can open.
Because of the increased fluid-loss rate, this behavior commonly results in a
screenout, often within minutes after such behavior is observed.
In general, type III behavior will be followed by undesired behavior such as
excessive fracture height growth or a screenout. For this reason, the net pressure
where the flat slope occurs is termed the critical pressure. For height growth, critical
pressure is approximately 70 to 80% of the closure stress difference between the
zone where the fracture was initiated and the confining formations. For natural
fissures, critical pressure is about equal to the closure stress holding the fissures
closed. In any case, the value of critical pressure is itself valuable data. In most
cases, the critical pressure found from several wells in a field is reasonably constant.
Therefore, the critical pressure found for one well may be extrapolated to offset
wells, and for designing future treatments, efforts can be made to keep net pressure
below this critical pressure. If this is not possible, it may be possible to incorporate

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 5 of 23

other factors into the treatment, assuring a greater chance of achieving the desired
fracture penetration without a screenout.
Several techniques may be used to reduce fluid loss. Refer to Appendix E  Fluid
Loss for information.
Type III Behavior
Predictions can be made concerning what will happen if one of the other variables
decreases by referring back to the continuity equation. For example, during type I
behavior, fracture length increases with time (∆L > 0). If something occurs so that
length extension stops, then one or both of the other variables, ∆p or ∆h must
increase (that is, pressure must increase faster or fracture height must grow). If the
fracture is well contained, then pressure must increase. This behavior is termed type
III behavior.
Refer to Fig 1. Type III behavior, with a slope of 1 indicates restricted extension at
the tip. The most common occurrence of this is when the pad volume has become
depleted and proppant reaches the fracture tip, arresting extension. This is called a
tip screenout.
A proppant bridge resulting from slurry dehydration due to natural fissures or height
growth will also cause type III behavior. When a fracture grows out of zone into a
formation with greater closure stress, the fracture “pinches down” at the boundary
between the zones. This width restriction can cause proppant to bridge, but allow
fluid to pass, dehydrating the slurry remaining in the main fracture. Once the slurry
dehydrates sufficiently, a plug is formed. The plug cannot be moved down the
fracture.
In addition to proppant bridging, restricted extension and type III behavior can occur
from the buildup of excessive fluid-loss additives in the pad. This causes a
restriction at the tip and penetration into a higher stress region because of pore
pressure gradients from prior production or because of lithology changes (that is,
limited dimensions of lenticular or channel sands). Calibration treatments are useful
to identify these causes for restricting fracture penetration.
Refer to Fig. 1. A slope greater than 1 indicates that the restriction is closer to the
wellbore. For instance, a wellbore screenout shows up as a near vertical line on the
plot. Restrictions farther from the well will have slopes closer to 1.
The approximate distance to the bridge can be calculated using Eq. 1.

qE
dr =
h 2∆ p / ∆ t (1)
Where:
dr = distance to restriction (ft)
q = pump rate (BPM)

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 6 of 23

E = Modulus of elasticity, Young's Modulus


h = fracture height (ft)
∆p /∆t = rate of pressure increase (psi/min).
Eq. 1 assumes a constant fracture height, that the bridge is stationary, and
calculates the maximum distance to the restriction. This calculation can then be
used for designing future fracture treatments. A near-wellbore bridge is likely due to
natural fissures, height growth or a slug of high concentration or proppant polymer.
A bridge at the tip may be caused by pad depletion, improper fracture design or fluid
problems.
Type IV Behavior
Refer to Fig. 1. Type IV behavior, a negative slope, is an indication of rapid,
unrestricted height growth. The radial-type fracture exhibits this response as well as
a fracture that breaks through any confining barriers.
If the fracture grows vertically through confining barriers into a zone of lower closure
stress, unrestricted height growth will occur. The relatively overpressured fracture
entering the low stress zone grows rapidly into the zone, increasing fracture volume
and fluid loss. As the pressure decreases, the fracture in the intermediate zone may
close. However, proppant bridging at the boundary of the main zone and the
confining barrier causes slurry dehydration in the main fracture. Unrestricted height
growth accelerates the process and a screenout results soon after, even at low
proppant concentrations.
The slurry dehydration, decreasing width and height-growth can be significantly
reduced by placing a relatively impermeable mixture of fluid-loss additives (100 to
300 mesh) between the pad and the proppant stages. The DIVERTAFRAC* Service,
INVERTAFRAC* Service, or both services can also be used. These techniques will
form an impermeable bridge at the pinch-point.
Foam fracturing fluids can also retard height-growth because of the yield stress for
these fluids. The yield stress permits a pressure gradient to be developed which is
proportional to the yield stress and inversely proportional to the width which is very
small at the pinch-point.
Type IV behavior may also be seen from the beginning of a fracture treatment. This
indicates no height-confinement, and that the fracture is expanding rapidly. This is
not as catastrophic as declining net pressure late in the treatment. However, future
fracture designs should incorporate a fracture model (KGD, radial) for an
unconfined-height fracture.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 7 of 23

Fig. 1. Slope interpretation for the Nolte-Smith plot.

2.1.2 In-Situ Stress Requirements


The in-situ stress requirements for confined height-growth and to keep natural
fissures from opening are very significant. The example in Fig. 2 indicates for a net
pressure of 1500 psi, the stress contrast in the vertical direction must exceed
1500 psi and in the horizontal direction must exceed 1200 psi. Stress contrasts of
these magnitudes cannot be assumed to occur for most reservoirs. Therefore,
effective fracturing using normal practices cannot be assumed to occur for most
reservoirs.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 8 of 23

Fig. 2. Example of required stress contrasts.

2.1.3 References
Comprehensive discussions of injection pressure analysis techniques are provided in
the following literature.
Reservoir Stimulation, Chapter 7, Economides, M.J., and Nolte, K.G. (eds.),
Schlumberger Educational Services (1989).
A Practical Companion to Reservoir Stimulation, Chapter D, Economides, M.J.,
Schlumberger Educational Services (1991).
Recent Advances in Hydraulic Fracturing, Chapter 14, Gidley, J.L., Holditch, S.A.,
Nierode, D.E., Veatch, R.W. (eds.), SPE Monograph Volume 12, Society of
Petroleum Engineers (1989).
Ayoub, J.A., Brown, J.E., Barree, R.D. and Elphick, J.: “Diagnosis and Evaluation of
Fracturing Treatments,” paper SPE 20581.

2.2 Pressure Decline Analysis


The analysis of the fracture after pumping allows the characterization of fracture
geometry and the determination of the closure pressure, leakoff coefficient, and
fracturing fluid efficiency.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 9 of 23

2.2.1 References
Comprehensive discussions of pressure decline analysis techniques are provided in
the following literature:
Reservoir Stimulation, Chapter 7, Economides, M.J., and Nolte, K.G. (eds.),
Schlumberger Educational Services (1989).
A Practical Companion to Reservoir Stimulation, Chapter D, Economides, M.J.,
Schlumberger Educational Services (1991).
Recent Advances in Hydraulic Fracturing, Chapter 14, Gidley, J.L., Holditch, S.A.,
Nierode, D.E., Veatch, R.W. (eds.), SPE Monograph Volume 12, Society of
Petroleum Engineers (1989).

2.3 Fracture Height Prediction and Post-Treatment Measurements


Fracture height prediction is based on the prediction and measurement of rock
properties in the layer above and below the target zone. These properties, properly
treated, result in the prediction of the vertical propagation of the fracture. Post-
treatment measurements including temperature logs, radioactive logs (in conjunction
with radioactive materials), and sonic logs allow the estimation of the fracture height.

2.3.1 Sonic Logs


Sonic waveforms are sensitive to the presence of fractures, even in cased-hole. In
hard formations, conventional monopole sonic tools such as the Schlumberger
Array-Sonic* fullwave sonic velocity tool or even the Long-Spaced Sonic tool will
measure a shear component and can be used to detect fractures. In soft formations,
the shear component is absent from monopole sonic waveforms and a dipole sonic
tool (Schlumberger DSI* tool) is required to measure the shear wavetrain.
The technique consists of acquiring a sonic waveform VDL before the fracturing
treatment and another after the fracturing treatment. The difference is usually very
clear. In obvious cases, a single pass may be sufficient. One drawback is the
limited depth of investigation (usually around a foot). Consult Wireline engineers to
determine the tool best suited for the area.

2.3.2 References
Comprehensive discussions of fracture height-prediction techniques are provided in
the following literature.
Reservoir Stimulation, Chapter 10, Economides, M.J., and Nolte, K.G. (eds.),
Schlumberger Educational Services (1989).
A Practical Companion to Reservoir Stimulation, Chapter B, Economides, M.J.,
Schlumberger Educational Services (1991).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 10 of 23

3 Treatment Performance Monitoring


Treatment performance monitoring involves data acquisition, and permits estimation
of the results of the fracturing treatment in its geologic setting. There are two
purposes for treatment performance monitoring.
1. The first purpose is to assess the fracture properties (length and conductivity)
achieved in the treatment against those prescribed initially by the client or
derived in the design analysis.
2. The second purpose is to provide the data which permits determination of the
full-scale mechanical and permeability properties of the host medium for the
treatment.
Fracture data is acquired during and immediately following the treatment execution
and during well production and pressure build-up tests. Because the data
acquisition and subsequent analysis are essentially independent for the two separate
monitoring exercises, the opportunity is provided for checking the consistency of the
models, assumptions and input data used separately in fracture engineering and
reservoir engineering for the well. Noting that in each exercise, the data set is
under-determined, (that is, there is less data than is needed to provide a unique
identification of the parameter set describing the system), some effort may be
required to reconcile the different models, the results obtained from them and the
analytical procedures used in the two fields of practice.
The time history of bottomhole treating pressure is the primary data set of the
treatment data. Because different aspects of formation mechanical behavior are
expressed in different stages of a treatment, a complete pressure history must be
collected. For example, early in a treatment, fracture growth may involve transition
from a radial to a confined fracture. That part of the pressure record contains unique
information on formation height, providing an essential geometric scale in the data
set. After shut-in, the pressure decline record contains information on stress barriers
and the leak-off coefficient. Ideally, other data should be collected which provide a
capacity to identify unique components of system performance. In a MultiFRAC*
treatment, cross-flow between layers after shut-in is related directly to the
compliance of the various layers.
In comparison with the data needs for thorough analysis of pressure and other
treatment data, those for well production and pressure build-up analysis have been
well defined through consideration of the principles of inverse analysis. A
comparable effort is required in treatment data set definition for treatment record
analysis.

3.1 Inverse Analysis of Treatment and Production Data Records


Inverse analysis is the technique by which a physical system is characterized from
observations of its response to imposed perturbations (disturbances), using a
simulator which represents the detailed behavior of the system. The technique was
developed initially to characterize the thermomechanical properties of bodies under

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 11 of 23

extreme loading conditions, such as rocket nose cones under re-entry conditions. In
its more general application, a system simulator and a search routine (or polytope)
may be used to identify the values of a parameter set which minimize an error
residual (or objective function) in a multi-variant parameter space. In structural
mechanics, for example, the method is used to invert vibration records of frames to
obtain the stiffness of individual members and of joints. In reservoir engineering, a
pressure build-up record is inverted using a reservoir simulator and a suitable
polytope to define formation properties of interest. The ZODIAC* software is an
inverse analysis application using a suite of simulators driven by a polytope called
“CONREG”.
In the application of inverse analysis techniques to a treatment record, the purposes
are to identify fracture length and conductivity, review the adequacy of the postulated
setting model and recover improved data on the in-situ mechanical properties of the
formation and its setting. If suitable constraints can be provided as inputs to the
analysis, it may return a more reliable modelization of the medium, particularly on
features not previously accounted for in the model. Estimates of formation
mechanical properties, for example, may reflect the average conditions which
operate on the in-situ scale rather than those obtained from non-representative
laboratory tests on disturbed specimens. The output from the inverse analysis
provides an improved set of setting characterization data, and may lead to significant
modification of the setting modelization, formation constitutive model and fracture
simulation procedures used in subsequent designs of treatments in adjacent wells in
the formation.

3.1.1 Fracture Characterization Using the ZODIAC Software


The ZODIAC software contains a general purpose pressure transient analysis model
that can be used to interpret the transient behavior of unfractured, horizontal, and
vertically fractured wells. Pressure transient testing of oil and gas wells is
considered by most reservoir engineers to be one of the more reliable means of
characterizing the production characteristics of a reservoir. Pressure transient tests
of unfractured wells can be used to obtain estimates of the average conductivity of a
reservoir, any near well flow impairment and interference with adjacent wells.
Pressure transient testing is also commonly used to evaluate the effectiveness of
hydraulic fracturing treatments. A significant limitation of some of the earlier
fractured well pressure transient analysis models was that there were so many
limiting assumptions used in the development of the interpretation models that the
models often bore little resemblance to the types and shapes of vertical fractures
that were actually being created. This limitation, together with the limitations of the
models to consider the practical, non-ideal reservoir characteristics such as
permeability, anisotropy and boundary effects, and the fracture properties of fracture
face skin and storage, often resulted in the under-estimation of the actual fracture
lengths.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 12 of 23

The new fractured well interpretation models that are being added to the ZODIAC
software will provide a means of obtaining more realistic estimates of the propped
fracture dimensions and conductivity than has been possible previously. The
practical analysis features that have been added to the ZODIAC software that can be
used to more properly interpret the pressure transient behavior of vertically fractured
wells includes:
• the effects of spatially-variable fracture height
• fracture width effects
• the effect of conductivity distributions
• the effects of fracture face skin damage
• fracture storage effects
• the effects of reservoir permeability anisotropy
• finite reservoir boundary effects.
The analysis features of fracture storage and boundary effects are not operational in
the current version of the ZODIAC software but have been fully developed and will
be available in the next engineering code release.
The pressure transient interpretation models in the ZODIAC software can be used to
detect and generally quantify each of these reservoir and fracture characteristics.
The summary that follows addresses each of the new features that are available in
the ZODIAC software to obtain a better interpretation of the pressure transient
behavior of vertically fractured wells and to more properly characterize the created
fracture dimensions and conductivity.

3.1.1.1 Fracture Storage


The general pressure transient analysis models that have been developed for the
ZODIAC software include fracture flow models that consider (or neglect) the storage
effects of the vertical fracture due to the contrast in the hydraulic diffusivities of the
reservoir and the propped fracture. The effects of fracture storage are minimal for
moderate- to high-dimensionless-conductivity fractures at early time; the effects of
fracture storage on the pressure transient behavior can be observed for much of the
duration of the transient test period.
When referring to the fracture storage effect, it is most convenient to use the
dimensionless hydraulic diffusivity to quantify the pressure transient behavior of the
system, since the dimensionless hydraulic diffusivity and the dimensionless fracture-
conductivity-height product are the only two fracture parameter groups that are
necessary to parameterize the pressure distribution in the fracture as a function of
time and space. The fracture pressure distribution is of course a function of the
reservoir properties, which are related to the fracture flow solution through the
fracture flux distribution. A comparison of the pressure transient behavior of a high-
conductivity fracture (CfD = 300) with the finite-conductivity fracture models that

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 13 of 23

neglect or consider the fracture storage effects is shown in Fig. 3. A similar


comparison is shown in Fig. 4 for a low-conductivity fracture (CfD = 3). The fracture
storage has minimal effect on the pressure transient behavior of the low-conductivity
fracture after a dimensionless time of about 10-3. However, the pressure transient
behavior of the high-conductivity fracture is observed to be significantly affected by
the contrast in the hydraulic diffusivities of the reservoir and the fracture as late a
dimensionless time of 10-2. Using a fracture interpretation model that neglects the
fracture storage effects will tend to produce estimates of fracture conductivity that
are too high and will result in under-estimation of the actual fracture length.

3.1.1.2 Fracture Face Skin Damage


The current version of the ZODIAC software can be used to quantify fracture face
skin damage effects on the pressure transient behavior of a finite conductivity
fracture using a model that considers the fracture storage effects to be negligible.
The next engineering code release of the ZODIAC software will include a finite-
conductivity fracture interpretation model that considers the fracture storage effects
of the system. Fig. 5 shows the effect of fracture face skin damage on the pressure
transient behavior of a moderate-conductivity fracture (CfD = 10) with fracture
storage. The corresponding comparison of the fracture face skin damage effects on
a finite-conductivity fracture (same dimensionless conductivity; fracture storage
effects considered to be negligible), is shown in Fig. 6. A similar set of comparisons
are shown in Fig. 7 and Fig. 8 for a low-conductivity (CfD = 1) fracture. These
comparisons illustrate that each of these factors (fracture face skin and storage) can
significantly affect the interpretation of the pressure transient behavior of finite-
conductivity fractures, both individually and in combination with each other.

3.1.1.3 Variable Fracture Conductivity


The pressure transient behavior of a finite-conductivity fracture with spatially varying
fracture conductivity is shown in Fig. 9 for three fracture conductivity distributions, in
which, the fracture has high average dimensionless conductivity. Little noticeable
effect is observed for the spatially variable fracture conductivity for the practical
range of transient test times.
The fracture conductivity distributions used in this comparison were uniform, linearly
varying, step profile dimensionless fracture conductivity distributions. The
dimensionless fracture conductivity for the variable conductivity distribution schemes
were specified as equal to 500 at the wellbore and 0 at the fracture tip. The step
profile used in this comparison consisted of 50% reductions in the dimensionless
fracture conductivity at each quarter of the fracture length. The corresponding
dimensionless fracture conductivity distribution that was used in this comparison for
the step profile was CfD = 500 for the first quarter of the fracture length, CfD = 250 for
the second quarter, CfD = 125 for the third quarter, and a CfD = 62.5 for the last
quarter of the fracture length (nearest the fracture tip).

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 14 of 23

A similar comparison is shown in Fig. 10 for a finite-conductivity fracture with a


moderate average fracture conductivity (CfD). A similar set of conductivity profiles
were used in this case except that the specified dimensionless fracture conductivity
equaled 10. In this case, the effects of a spatially variable fracture conductivity
distribution are more significant over the range of transient times of interest. For
even lower average dimensionless conductivity fractures, the effect of the spatial
variation in the fracture conductivity becomes more pronounced.

3.1.1.4 Reservoir Permeability Anisotropy


Fig. 11 and Fig. 12 show the dimensionless and dimensional pressure transient and
derivative behaviors of a finite-conductivity fracture with a uniform fracture
conductivity of 400,000 md-ft in an anisotropic reservoir, with various levels of
reservoir permeability anisotropy. The fracture and reservoir parameters used
(xf = 400 ft, kmax = 1.0 md) correspond to a dimensionless fracture conductivity of 1000
in an isotropic reservoir with kx = ky = 1.0 md. Fig. 13 and Fig. 14 show the
dimensionless and dimensional pressure transient behaviors of a finite-conductivity
fracture with uniform fracture conductivity of 2000 md-ft. The same reservoir and
fracture parameters are used in this case as in the previous example. In these
examples, the dimensionless fracture conductivity for the isotropic reservoir case is
equal to 5.0. The anisotropic reservoir model is available in the current release of
the ZODIAC software.

3.1.1.5 Finite Reservoirs


The next engineering code development release of the ZODIAC software will have
several finite reservoir solutions available for fractured wells. The finite reservoir
solutions that will be available for the analysis of vertically fractured wells in finite
reservoirs will be the models for a fractured well centered in a closed or constant
pressure circle, and a fractured well that can be off-centered in a closed rectangle.
Examples of the types of finite reservoir solutions that will be available in the next
release of the ZODIAC software are shown in Fig. 15, Fig. 16, and Fig. 17.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 15 of 23

Fig. 3. High-conductivity fracture comparison.

Fig. 4. Low-conductivity fracture comparison.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 16 of 23

Fig. 5. Fracture face skin damage, moderate-conductivity fracture.

Fig. 6. Fracture face skin damage, finite-conductivity fracture.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 17 of 23

Fig. 7. Fracture face skin damage comparison, low-conducitivy fracture.

Fig. 8. Fracture skin damage comparison, low-conductivity fracture.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 18 of 23

Fig. 9. Finite-conductivity fracturecomparison, high average dimensionless conductivity.

Fig. 10. Finite-conductivity comparison, moderate average fracture conductivity.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 19 of 23

Fig. 11. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft.

Fig. 12. Finite-conductivity comparison, uniform fracture conductivity 4000,000 md-ft.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 20 of 23

Fig. 13. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft.

Fig. 14. Finite-conductivity comparison, uniform fracture conductivity 2000 md-ft.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 21 of 23

Fig. 15. ZODIAC software examples.

Fig. 16. ZODIAC software examples.

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Treatment Evaluation Dowell
Page 22 of 23

Fig. 17. ZODIAC software examples.

4 Production Evaluation
The inference of fracture length from the reservoir response is very sensitive to
uncertainties such as inferred skin, formation permeability, and fracture conductivity.
For a 5% change in skin, the minimum length change is 30%. The ability to infer a
representative length diminishes rapidly for a dimensionless fracture conductivity
less than two. More significant are the limitations of constant, homogeneous, and
isotropic reservoir permeability and homogeneous fracture conductivity. These
limitations produce inferred lengths which can be substantially less than the actual
length. Further, these effects are cumulative and can result in the inferred length
being only a small fraction of the actual case, or more importantly, the production
much less than anticipated and a false indication of treatment failure. Except for
anisotropic permeability, all the limitations can be diagnosed or quantified by
standard field measurements and analyses. Special testing of oriented core can
identify matrix-permeability anisotropy and provide a relatively accurate assessment
of permeability variations with depth.
Reservoir analysis typically assumes constant, isotropic, and homogeneous
conditions for the reservoir and fracture conductivity. Heterogeneous reservoir and
conductivity effects will cause the well production to be less than expected from the

DOWELL CONFIDENTIAL
Section 600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Treatment Evaluation
Page 23 of 23

assumption of homogeneous properties. Heterogeneous reservoir conditions may


give a false indication of fracturing treatment failure.
In the case of constant, isotropic, and homogeneous conditions, evaluation can be
routinely applied by district level engineers with standard training and CADE
functionality. The evaluation of well performance is a comparison of predicted
versus actual performance.
The FORECAST module in the FracCADE* software may be used for production
prediction. User information for the FORECAST module is provided in the
FracCADE Users Manual.
In the case of heterogeneous reservoir and conductivity effects, accurate evaluation
will require proficient use of the most advanced techniques and will require specially
trained experts within the area or region. The evaluation may require several weeks
to months for completion.

4.1 References
Comprehensive discussions of production analysis techniques are provided in the
following literature.
Reservoir Stimulation, Chapter 11, Economides, M.J., and Nolte, K.G. (eds.),
Schlumberger Educational Services (1989).
A Practical Companion to Reservoir Stimulation, Chapter F, Economides, M.J.,
Schlumberger Educational Services (1991).
Recent Advances in Hydraulic Fracturing, Chapter 15 and Appendix K, Gidley, J.L.,
Holditch, S.A., Nierode, D.E., Veatch, R.W. (eds.), SPE Monograph Volume 12,
Society of Petroleum Engineers (1989).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 1 of 81

DataFRAC SERVICE

1 Introductory Summary............................................................................................................. 6
1.1 Closure Test......................................................................................................................... 7
1.1.1 Closure Test in a Permeable Zone ............................................................................. 7
1.1.2 Closure Test in a Nonpermeable Zone....................................................................... 9
1.2 Calibration Test.................................................................................................................... 9
1.3 Applications........................................................................................................................ 10

2 Design ..................................................................................................................................... 11
2.1 Preparatory Engineering .................................................................................................... 11
2.1.1 Breakdown/Diversion Treatment .............................................................................. 11
2.1.2 Preliminary Fracture Design ..................................................................................... 11
2.1.3 Fracture Height......................................................................................................... 11
2.1.4 Wellbore Logging...................................................................................................... 12
2.1.4.1 Temperature and Gamma-Ray Logs ............................................................. 12
2.1.4.2 Fracture-Height Logs ..................................................................................... 13
2.1.5 Perforating ................................................................................................................ 13
2.1.5.1 Wellbore Restrictions ..................................................................................... 13
2.1.5.2 Perforation Phasing ....................................................................................... 14
2.1.5.3 Perforation Size ............................................................................................. 14
2.2 Closure Test....................................................................................................................... 15
2.2.1 Fluid Selection .......................................................................................................... 15
2.2.2 Injection Rates and Number of Steps ....................................................................... 15
2.2.3 Step Duration............................................................................................................ 15
2.2.4 Flow-Back Rate ........................................................................................................ 16
2.3 Calibration Test.................................................................................................................. 17
2.3.1 Fluid Selection .......................................................................................................... 17
2.3.1.1 Foam.............................................................................................................. 17
2.3.2 Fluid Volume............................................................................................................. 17
2.3.3 Fluid Break-Time ...................................................................................................... 18

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 2 of 81

2.3.4 Fluid-Loss Additives ..................................................................................................18


2.3.5 Duration of Pressure Decline ....................................................................................18
2.4 Special Considerations in the DataFRAC Design...............................................................18
2.4.1 The Influence of Wellbore Fluid ................................................................................18
2.4.2 Prepad.......................................................................................................................18
2.4.3 Closure Pressure less than Hydrostatic Pressure.....................................................19
2.4.4 Post-Job Wireline Surveys ........................................................................................19
2.5 Terminology........................................................................................................................19
2.5.1 Fracture Extension Pressure.....................................................................................19
2.5.2 Initial Shut-in Pressure ..............................................................................................19
2.5.3 Closure Pressure ......................................................................................................19
2.5.4 Rebound Pressure ....................................................................................................19
2.6 Equipment Requirements ...................................................................................................20
2.6.1 Monitoring Equipment ...............................................................................................20
2.6.2 Pumping Equipment..................................................................................................20
2.6.3 Pressure Measuring Equipment................................................................................20
2.6.3.1 Surface Measurement Methods .....................................................................20
2.6.3.2 Bottomhole Pressure Gauge Measurement ...................................................22
2.6.4 Treating Equipment...................................................................................................23
2.6.5 Flowback Equipment.................................................................................................23
2.6.5.1 Magnetic Flowmeters .....................................................................................23
2.6.5.2 Turbine Flowmeters........................................................................................23
2.6.5.3 Chokes and Gate Valves................................................................................23

3 Execution ................................................................................................................................24
3.1 Pre-Performance Guidelines ..............................................................................................24
3.2 Closure Test .......................................................................................................................27
3.2.1 Step-Rate Phase.......................................................................................................27
3.2.2 Flowback Phase........................................................................................................32
3.2.2.1 Flow Control ...................................................................................................32
3.2.2.2 Flowmeters.....................................................................................................34
3.2.3 Closure Test Modifications........................................................................................34

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 3 of 81

3.3 Calibration Test.................................................................................................................. 35


3.3.1 Injection Phase ......................................................................................................... 35
3.3.2 Pressure-Decline Phase ........................................................................................... 36
3.3.3 Contingency Plans.................................................................................................... 36

4 Evaluation............................................................................................................................... 36
4.1 Closure Test Analysis ........................................................................................................ 37
4.1.1 Step Rate  The BHP-Versus-Rate Plot ................................................................. 37
4.1.2 Flowback  The BHP-Versus-Time Plot.................................................................. 37
4.1.3 Confirmation of Closure Pressure............................................................................. 38
4.1.4 Rebound Pressure.................................................................................................... 40
4.2 Calibration Injection for Fracture Geometry ....................................................................... 40
4.2.1 Elastic Fracture Compliance..................................................................................... 41
4.2.2 Pressure During Pumping......................................................................................... 43
4.2.2.1 Fluid Flow and Pressure in Fracture .............................................................. 43
4.2.2.2 Nolte-Smith Plot and Evolution of Pressure During Pumping ........................ 45
4.2.3 Deviations from Ideal Geometry ............................................................................... 46
4.2.3.1 Height Growth ................................................................................................ 46
4.2.3.2 Fissures ......................................................................................................... 47
4.2.3.3 T-Shape Fracture........................................................................................... 48
4.2.4 Pressure Capacity .................................................................................................... 49
4.2.5 Near-Wellbore Restriction......................................................................................... 50
4.2.6 Fracturing Pressure Interpretation Summary ........................................................... 53
4.2.6.1 Example of Radial Fracture ........................................................................... 54
4.2.6.2 Simulation of Pressure During Pumping and Decline .................................... 54
4.3 Calibration Decline for Fluid-Loss Behavior ....................................................................... 56
4.3.1 Review of Decline Analysis....................................................................................... 56
4.3.2 Volume Function g.................................................................................................... 58
4.3.3 Fluid Efficiency.......................................................................................................... 59
4.3.4 Decline Function G ................................................................................................... 61
4.3.5 Non-Ideal Behavior ................................................................................................... 64
4.3.5.1 Change in Fracture Penetration After Shut-in................................................ 64

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 4 of 81

4.3.5.2 Height Growth ................................................................................................65


4.3.5.3 Pressure-Dependent Leakoff .........................................................................66
4.3.5.4 Spurt...............................................................................................................69
4.3.5.5 Closure Pressure Change ..............................................................................69
4.3.5.6 Compressible Fluids.......................................................................................71
4.3.6 Fluid Efficiency Based on Pressure Analysis ............................................................72
4.3.7 Decline-Analysis Procedure ......................................................................................73
4.3.8 Steps to Correct Decline Analysis Using the FracCADE Software............................75
4.3.8.1 The DataFRAC Software................................................................................76
4.3.8.2 G-plot Interpretation by the DataFRAC Software ...........................................77
4.3.8.3 Modulus, Height or Fracture Toughness Calibrations ....................................77
4.3.8.4 The β Ratio.....................................................................................................78
4.3.9 Post Proppant Fracture Analysis...............................................................................80
4.3.10 References..............................................................................................................81
FIGURES
Fig. 1. The effect of proppant-pack damage and fracture length on fracture NPV. ......................6
Fig. 2. Fracture extension pressure (unequal time steps). ...........................................................7
Fig. 3. The typical closure test......................................................................................................8
Fig. 4. The G-plot (idealized). .....................................................................................................10
Fig. 5. Channel restriction at the wellbore. .................................................................................13
Fig. 6. The relation of perforation diameter and proppant concentration. ..................................14
Fig. 7. The effects of differing flowback rates. ............................................................................16
Fig. 8. The change in surface pressure during closure in deep, hot wells..................................21
Fig. 9. Hydrostatic head changes during closure. ......................................................................22
Fig. 10. The DataFRAC Service rig-up when pumping conductive fluids. ..................................25
Fig. 11. The DataFRAC Service rig-up when pumping nonconductive fluids. ............................26
Fig. 12. Friction pressure of water in the tubing and casing. ......................................................28
Fig. 13. Friction pressure of water in the annulus.......................................................................29
Fig. 14. Friction pressure of brine in the tubing and casing........................................................29
Fig. 15. Friction pressure of brine in the annulus. ......................................................................30
Fig. 16. Friction pressure of diesel in the tubing and casing. .....................................................30
Fig. 17. Friction pressure of diesel in the annulus. .....................................................................31
Fig. 18. Flow rate versus differential pressure in perforations....................................................31
Fig. 19. Flowback test (after Nolte, 1982/1994)..........................................................................38
Fig. 20. Effect of closure on BHP versus square root of t and G- plots. .....................................39
Fig. 21. Rebound pressure; lower bound of closure pressure....................................................40
Fig. 22. Analogy of a pressurized crack to a pre-loaded spring. ................................................42
Fig. 23. Evolution of fracture geometry and pressure during pumping.......................................45
Fig. 24. Pressure and width for height growth through barriers (after Nolte, 1989)...................46

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 5 of 81

Fig. 25. Pressure and width for opening natural fissures (after Nolte, 1989)............................ 47
Fig. 26. Pressure and width for T-shape fracture (after Nolte, 1989)........................................ 48
Fig. 27. Definition of pressure capacity from in-situ stresses..................................................... 50
Fig. 28. Stress state within the entrance of deviated well or stress. .......................................... 51
Fig. 29. Mohr circle of deviated well or stress. ........................................................................... 52
Fig. 30. Nolte-Smith plot of fracturing pressure. ........................................................................ 53
Fig. 31. Net pressure with radial fracture (after Smith et al. 1987). .......................................... 54
Fig. 32. Measured and simulated net pressure: opening natural fissures (after Nolte, 1982). . 55
Fig. 33. Example of fracturing-related pressures (after Nolte, 1982). ........................................ 56
Fig. 34. Schematic for fracture area and time............................................................................ 57
Fig. 35. Dimensionless volume function for fracture closure (after Nolte, 1986)....................... 59
Fig. 36. Efficiency from closure time for no proppant, no spurt loss during pumping and
other ideal assumptions given in Section 4.3.1 (after Nolte, 1986). ............................. 60
Fig. 37. Conceptual response of pressure decline versus Nolte time-function
(after Castillo, 1987). .................................................................................................... 62
Fig. 38. Penetration change during shut-in (after Nolte, 1990). ................................................. 65
Fig. 39. Diagnostic for height growth from decline data (after Nolte, 1990)............................... 66
Fig. 40. Diagnostic for stress sensitive fissures from injection and decline (after Nolte, 1990). 67
Fig. 41. Decline analysis for filtrate and reservoir control leakoff (after Nolte, 1993)................ 68
Fig. 42. Stress change during injection/shut-in for Cc (after Nolte et. al., 1993)......................... 70
Fig. 43. Relative volume change of gas (after Nolte et. al., 1993). ........................................... 72
Fig. 44. Decline analysis using “¾” rule (after Nolte, 1990). ...................................................... 74
Fig. 45. Pressure and flow rate in fracture before and after shut-in (after Nolte, 1986)............ 79
Fig. 46. Diagnostic for closing on proppant from decline data (after Nolte, 1990). ................... 80
TABLES
Table 1. Approximate Choke Settings For Flowback Of Oil-Base Fluids (Sg = 0.7) .................. 33
Table 2. Approximate Choke Settings for Flowback of Water-Base Fluids (Sg = 1.0)............... 34
Table 3. Interpolated Values of α Over the Full Range of n....................................................... 58
Table 4. Values of Decline Function "G" .................................................................................... 63
Table 5. Correction Factors f c As Function Of ∆tD ...................................................................... 75

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 6 of 81

1 Introductory Summary
The DataFRAC* Service determines the in-situ parameters critical to optimum
fracture treatment design. These parameters are specific to each formation and
often to each well. Assumed or inaccurate parameter values can result in the
following.
• Premature screenout and reduced fracture penetration caused by pad fluid
depletion.
• Unpropped fracture, increased damage to proppant-pack conductivity and
increased treatment cost because of excessive pad volume.
Both outcomes result in reduced net present value (NPV), illustrated in Fig. 1.

Fig. 1. The effect of proppant-pack damage and fracture length on fracture NPV.
(THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY.)
The DataFRAC Service typically consists of two tests  a closure test and a
calibration test.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 7 of 81

1.1 Closure Test


The closure test determines closure pressure  the minimum in-situ rock stress.
Accurate determination of closure pressure is important because all fracture analysis
is referenced from it. Closure pressure is also used for proppant selection.
The closure test is recommended as one of the initial procedures of any field
stimulation operation. Performance of a valid closure test
• ensures the zone has been fractured (a necessary condition for valid
performance of other tests)
• provides upper and lower bounds for determination of the closure pressure
• defines the required range of pump rates for extending a fracture in the zone.

1.1.1 Closure Test in a Permeable Zone


The closure test in a permeable zone is a step-rate/flowback procedure. A
Newtonian fluid is injected at an increasing rate until fracture extension occurs. A
pressure versus rate plot will show two distinct slopes, the intersection of which
indicates fracture extension pressure (Fig. 2). The change in slope in is a result of
the different pressure responses for matrix leakoff and fracture extension at the
higher rate. This pressure is normally 50 to 200 psi greater than closure pressure
because of fluid friction in the fracture and fracture toughness.

Fig. 2. Fracture extension pressure (unequal time steps).


(THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY.)

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 8 of 81

Another indication of fracture extension pressure comes from a bottomhole pressure


versus time plot and is illustrated in Fig. 3. The pressure steps above fracture
extension pressure have squared shoulders compared to the rounded shoulders
characteristic of matrix leakoff.

Fig. 3. The typical closure test.


(THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY.)
Pumping continues for five to ten minutes after fracture extension. The well is then
flowed-back at a constant rate. Flowback is started immediately after the final step
and is held constant until pressure has fallen to about 200 psi above the initial
wellbore pressure. The pressure response will show a distinct reversal in curvature
once closure has occurred (Fig. 3), indicating a change of fluid withdrawal from the
open fracture to withdrawal through the matrix. The rebound pressure after shut in
serves as a lower bound to closure pressure.
Perforation friction pressure is another important parameter that is determined from
the step-rate/flowback test. At shut-in, the immediate bottomhole pressure drop is
the pressure loss in the perforations during the last stage of the step-rate test. The
pressure loss will give an indication of potential wellbore problems, usually unopened
perforations. Reperforating should be considered if the pressure loss is
unacceptable.
The closure pressure is determined by quantitative analysis of bottomhole pressure
versus time using the Pressure Analysis and DataFRAC modules in the FracCADE*
software.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 9 of 81

The closure pressure may also be determined from a shut-in/decline test by analysis
of a square-root plot. However, the shut-in/decline method does not provide a
definitive indication of the closure pressure and is not the preferred method.

1.1.2 Closure Test in a Nonpermeable Zone


The closure test in a nonpermeable zone (shale) is an injection/shut-in procedure
where a small quantity (tens of gallons) of a Newtonian fluid is injected at low rate.
Pumping stops and an initial shut-in pressure is observed. Local stress is
approximately equal to the initial shut-in pressure; therefore, net pressure is
approximately equal to zero and the initial shut-in pressure is used to infer the stress.

1.2 Calibration Test


The calibration test is an injection/shut-in/decline procedure. A viscosified fluid
(without proppant) is pumped at proposed fracturing treatment rate. The well is then
shut in and a pressure decline analysis is performed.
The following critical design parameters are determined from the calibration test.
• fracture half-length (xf)
• fracture width (w)
• fracture height (hf)
• fluid-loss coefficient (C)
• Young's modulus (E)
• fluid efficiency (η).
The injection test determines the type of fracture being created; Perkins-Kern-
Nordgren (PKN), Khristianovic-Geertsma-de Klerk (KGD), or Geertsma-de Klerk
radial (RAD). Analysis of the net pressure versus time on a log-log scale (Nolte-
Smith plot) determines the type of model (PKN, KGD, or RAD) to use for decline
analysis. The injection test also serves as the pumping portion of the decline test.
Pressure decline after shut-in is monitored and is analyzed using the Pressure
Analysis, Decline Data and DataFRAC modules in the FracCADE software to
determine the parameters listed above.
The DataFRAC Service uses the G-plot for complete, consistent analysis. The
G-plot (illustrated in Fig. 4) replaces the curve-matching method and can accentuate
nonideal fracture behavior such as unrestrained height growth and extension after
shut-in and closure. Analysis results from the DataFRAC module in the FracCADE
software automatically update the fracture geometry simulator. The calculated net
pressure is compared and recorded with the net pressure observed at shut-in. This
dual analysis ensures a consistent set of parameters for the treatment design and
indicates potential nonideal fracture behavior when a pressure match cannot be
justified.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 10 of 81

Fig. 4. The G-plot (idealized).


(THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY)

1.3 Applications
The DataFRAC Service is an expense to the client that is not incurred if generally
available design data that is not specific to a particular well is used. However, this
service can increase the NPV when it results in optimization of a treatment design.
The DataFRAC Service can be routinely performed before all fracture treatments
when the objective is to optimize the treatment design and resulting production. It is
also an invaluable aid to assure the best possible treatment is performed in cases
where information is limited. Some opportunities where the DataFRAC Service
offers particular benefits are
• pilot projects or test wells that are critical to future development plans
• wells that are considered typical to a field where designs are being tested to
settle on an optimum
• exploration wells that have no history on which to design a treatment with a high
level of confidence
• areas where fracture response is not as anticipated and the cause requires
identification.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 11 of 81

2 Design

2.1 Preparatory Engineering


The DataFRAC Service is mainly analytical in nature. Other sources of data will
enhance the DataFRAC analysis.

2.1.1 Breakdown/Diversion Treatment


Perform a breakdown/diversion treatment (for example, acid ballout) prior to
performing a closure or calibration test to ensure that all perforations are open and
that the formation has been broken-down. The initial shut-in pressure recorded on a
breakdown/diversion treatment will be a very rough estimate of the closure pressure.

2.1.2 Preliminary Fracture Design


The parameters important to the DataFRAC Service are discussed below. Fracture
treatment design is provided in Treatment Design.
Before performing the DataFRAC Service, a fracturing treatment should be designed
using the best data available. Use the FracCADE software for the treatment design.
The fluid type, expected pad volume and efficiency, fracture geometry, and net
pressure will provide a reference for the same parameters that will be determined
from the DataFRAC analysis. A preliminary fracture design will also help to identify
unexpected or nonideal behavior during the closure and calibration tests.
If the preliminary fracture design indicates that the fracture capacity will be exceeded
(undesired height growth or opening of fissures), the DataFRAC Service will confirm
that and will quantify the fracture capacity based on actual, rather than assumed
pumping conditions. The subsequent fracture design can then be prepared with
either more confidence that the fracture capacity will not be exceeded or that special
techniques can be used to alleviate the problem.

2.1.3 Fracture Height


Fracture height affects fracture volume in two ways: directly, and through its effect on
width (determined by the fracture compliance). Accurate values for gross fracture-
height (formation gross height) and leakoff height (formation net thickness) are
critical to the DataFRAC analysis and to the ultimate success of the fracture design
and execution. If these values cannot be selected with a comfortable degree of
certainty prior to the fracture treatment, the need for the DataFRAC Service and
wireline surveys (logs) becomes even more critical for stimulation success.
The following methodology may be used to determine fracture height.
• Select “apparent” barriers from logs.
• Perform the DataFRAC Service to verify that height and Young's modulus match
with log-derived values.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 12 of 81

• Run pre- and post-job temperature logs or radioactive tracers and a gamma-ray
log (or all) to identify the actual fracture height.

2.1.4 Wellbore Logging


Pre- and postjob logs can give a starting point for height determination in the
analysis. Prior to performing the DataFRAC Service, request that the appropriate
wireline services be utilized to estimate fracture height (Gamma-Ray log, Sonic log),
leakoff height (SP log, Porosity log), and Young's modulus (Sonic log). Request
radioactive tracers for the calibration test. Request postfracture logs (Temperature
log and Gamma-Ray log) for fracture height verification.

2.1.4.1 Temperature and Gamma-Ray Logs


Temperature and gamma-ray logs are commonly used to determine fracture height.
Gross fracture-height is commonly determined from lithology information. Leakoff
height can be based on a porosity cut-off or gamma-ray/spontaneous potential (SP)
deflection. Normally, the height of any zone with greater than 1/3 deflection from the
shale base-line is considered leakoff height. Additional techniques to determine
fracture height are provided in Reservoir Stimulation.
During analysis, the following should be considered.
1. Logs only detect radioactive material and temperature differences a few inches
away from the wellbore.
2. The fracture tends to be away from the wellbore outside the perforated interval.
3. The formation must have both permeability and porosity to hold enough
radioactive fluid for detection.
In the first consideration, wellbore fracture height may not be the same as the
average height of the fracture because of deviated wellbore or zone, height growth
into the barriers at the wellbore or horizontal fractures. The net pressure (during
pumping) and a fracture simulator can give estimation of the average height. In the
DataFRAC module, height and Young's modulus are altered to make the Fracture
Geometry Sensitivity simulator (FGS) and the analysis (actual) net pressure match.
When the net pressures are matched, the heights and modulus should match with
those obtained from logs. If no match is obtained, then one of the sources may be
incorrect.
Shale barriers have very low permeability and porosity and will tend to “squeeze out”
any fluid during fracture closure. A more permeable and porous zone above the
shale will retain the fluid. A fracture may grow into this zone and the indication be
discounted because the shale barrier doesn't show radioactivity or temperature
change. This can also be missed if the wireline service company turns down the tool
sensitivity when away from the zone of interest.
Without an independent indication of fracture height, analysis is more difficult and
may be less accurate. Analysis will be enhanced with the aid of logs.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 13 of 81

2.1.4.2 Fracture-Height Logs


If a fracture-height log is available, use stress information from the log to limit net
pressure and, therefore, fracture height in the design. This can often demonstrate
the sensitivity of vertical fracture growth to pump rate and fluid rheology. Once
sensitivity is established, the need for the DataFRAC service is apparent to pinpoint
the critical design parameters and to calibrate the FGS simulator.

2.1.5 Perforating
Perforating technique can have a significant effect on the execution and evaluation
of the DataFRAC Service by affecting the breakdown and treating pressure.

2.1.5.1 Wellbore Restrictions


Wellbore restrictions will mask the formation pressure response while pumping. The
value for net pressure will be inaccurate because of a shift upward. Fracture model
selection may be affected. During the fracturing treatment the proppant will erode
the restrictions resulting in lower perforation friction pressure. A drop in perforation
friction pressure may be interpreted (falsely) as fracture height-growth.
Wellbore restrictions caused by improper or ineffective perforating techniques can
cause a screenout. Restrictions can cause the fracture to extend in an area apart
from the perforation tunnel, resulting in a significant increase in apparent perforation
friction pressure (Fig. 5).

Fig. 5. Channel restriction at the wellbore.


(THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY)

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 14 of 81

2.1.5.2 Perforation Phasing


Wells are commonly perforated with 0° phasing (perforations vertically aligned on
one side of the casing). For these cases, the orientation of the perforation with the
plane of the hydraulic fracture may be as large as 90°. With 0° phase perforations,
near perfect alignment will cause preferential propagation of one wing of the fracture
with very limited penetration of the companion wing. Channels are created and
cause higher treating pressures because of width restriction (Fig. 5).
Fig. 5 also shows a perforation that is approximately 30° out of phase to the fracture
plane (minimum stress). The fracturing fluid must partially circumvent the wellbore to
reach the fracture. Restrictions may develop, causing an increase in friction
pressure and creating the potential for proppant bridging. Even when a perforation
is directly in line with the fracture plane, the fracturing fluid must create a path
around the wellbore. With 90 or 120° phasing, the fracture plane will generally be
less than 30° from two perforations and will result in perforation access to both
fracture wings. (Note from Fig. 5 that 180° phasing would not alleviate the
misalignment).

2.1.5.3 Perforation Size


Fig. 6 illustrates the relation of perforation diameter and proppant concentration. A
perforation must be large enough to permit the proppant (at the maximum
concentration) to pass through and not bridge in the perforation.

Fig. 6. The relation of perforation diameter and proppant concentration.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 15 of 81

2.2 Closure Test


The general steps in a closure-test design are
1. Determine the fluid type.
2. Determine the injection rates and number of steps.
3. Determine the duration of steps.
4. Determine the flowback rate.
5. Determine equipment requirements.

2.2.1 Fluid Selection


In low-permeability formations, the closure test is usually performed with a
Newtonian fluid such as diesel or water containing 2% (wt:wt) potassium chloride. In
higher permeability formations (> 10 md) or in formations containing natural fissures,
viscosified fracturing fluids may be required to reduce the rate of fluid loss and
fracture closure during flowback. The same fluid as the pad fluid of the proposed
fracturing treatment would be a good choice in the case of high leakoff.

2.2.2 Injection Rates and Number of Steps


When injecting a Newtonian fluid, the range of rates is generally one to ten bbl/min
for larger and moderately permeable zones and approximately one-half these values
for smaller and very low permeability zones. After a breakdown/diversion treatment
has been performed, most zones (k > 0.01 md or h > 30 ft) will require a pump rate
greater than 3 bbl/min to exceed fracture extension pressure. The actual range for a
particular zone may require trial and error methodology; two or more attempts.
Ideally, three values of pressure (end of step) should fall below the extension
pressure to define the initial portion for flow into the matrix or a pre-existing fracture,
and a similar number of values above the extension pressure to define the portion
for extending the fracture. This allows the pressure versus rate plot to be drawn on
Cartesian coordinates using the last pressure before a rate change. The intersection
of the two straight lines (fracture extension pressure) provides an upper boundary for
closure pressure.
An additional step-rate/flowback test can be performed to verify correct closure. If
there were no pre-existing fracture, the plot of injection pressure versus bottomhole
pressure may show an overshoot of the extension pressure for one or two steps
because of the larger pressure required for breakdown and initiation of a fracture.

2.2.3 Step Duration


For the purpose of defining closure pressure, the duration of the individual rate steps
should be equal and can be relatively small. The time required for the pumping
equipment to change and maintain a constant rate (one or two minutes) is sufficient.
The last step is maintained for a longer time (five to ten minutes).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 16 of 81

All steps, except the last step, should be the same duration. The last step should be
long enough to establish some fracture volume, thus allowing the flowback, not
leakoff, to bring about closure. Five to ten minutes should be sufficient for the last
step.

2.2.4 Flow-Back Rate


The step-rate phase is followed by an immediate flowback at a constant rate.
Flowback should start immediately after shutdown. The rate must be held constant.
Flowback rate is controlled by an adjustable choke or a gate valve and is monitored
by a flowmeter. If the flowback rate is within the correct range, the resulting pressure
decline will show a characteristic reversal of curvature at the closure pressure. The
accelerated pressure decline at the curvature reversal is caused by the flow
restriction introduced when the fracture effectively closes. The correct range of
flowback rates must be determined by trial and error for any specific field; however,
the range is on the order of one-sixth to one-quarter of the fracture extension rate.
The effect of flow rates outside the correct range is shown in Fig. 7.
A second test may be required if the flowback rate made closure selection
impossible. The second test need not include a step-rate phase if clear fracture-
extension pressure was determined from the first test. Use a different rate the
second time.
Flowback until bottomhole pressure is within 200 psi of initial reservoir pressure. Do
not flow reservoir fluids into the wellbore by flowing back more than was injected. At
shut in, the pressure will rebound and stabilize.

Fig. 7. The effects of differing flowback rates.


THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 17 of 81

2.3 Calibration Test


The general steps in a calibration test design are
1. Determine the fluid type and injection rate.
2. Determine the fluid volume.
3. Determine the fluid break time.
4. Determine if fluid-loss additives are required.
5. Determine the pressure decline duration.

2.3.1 Fluid Selection


The type of fluid and injection rate for the calibration test are the same as the type of
fluid and injection rate of the proposed fracture treatment.

2.3.1.1 Foam
A foamed fluid may be used for the calibration test. However, the well must be
flushed with a linear fluid  a fluid containing no nitrogen, carbon dioxide or
crosslinker/activator. Gas in the flush volume will expand due to pressure decline
and temperature increase. This will cause fluid displacement into the fracture during
closure and will invalidate the decline analysis. If bottomhole pressure is calculated
from surface measurements, the hydrostatic pressure will change, adversely
affecting the calculations.

2.3.2 Fluid Volume


The fluid volume may be determined by using the FGS simulator in the FracCADE
software. Use the following methodology.
1. Determine the gross fracture-height and leakoff height.
2. Using a leakoff coefficient twice the value provided in the Fracturing Materials
Manual, calculate a minimum volume to ensure coverage of the zone if the KGD
or RAD model is selected (indicated by a lack of barriers). If the PKN model is
selected (indicated by significant barriers), calculate a volume sufficient to create
a fracture length greater than 1.5 times the fracture height.
3. If undesired height growth or fissure opening is suspected, treatment design
should incorporate methods to avoid them (DIVERTAFRAC* Service,
INVERTAFRAC* Service, or fluid-loss additives).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 18 of 81

2.3.3 Fluid Break-Time


Fluid break-time is designed for bottomhole static temperature and a long time
(compared with expected closure time). Five times the expected pumping time is a
good starting place.

2.3.4 Fluid-Loss Additives


FLA∗ 100 has particles large enough to be considered a proppant when used in a
calibration test. Therefore, FLA100 can cause a screenout and will affect analysis.
FLA100 is not recommended for use in a calibration test. However, in naturally-
fractured or high-leakoff formations, FLA100 can be used with caution if a sufficient
quantity of clean fluid is pumped ahead of it. Fluid-loss additive J84 or fluid-loss
additive J418 is not a screenout hazard and may be used in the entire fluid volume
for leakoff control.

2.3.5 Duration of Pressure Decline


The minimum time that pressure decline should be monitored is 1.25 times the
closure time or twice the injection time, whichever is longer. The closure time can be
estimated by using the Placement module in the FracCADE software. Estimate the
fluid and formation parameters and the volume of fluid to be pumped during the
calibration test. A very small proppant stage may be necessary to force the
Placement module simulator to run.

2.4 Special Considerations in the DataFRAC Design

2.4.1 The Influence of Wellbore Fluid


A large quantity of wellbore fluid injected prior to fracturing fluid entry can result in
substantial effects on analysis. If the static wellbore fluid volume is more than 10%
of the calibration test fluid volume, one of the following actions should be performed.
• Circulate the wellbore fluid out of the tubing with fracturing fluid.
• Bullhead the fracturing fluid to the top perforation at a low rate if circulation is not
possible. Allow the pressure to fall below closure pressure before starting the
calibration test.

2.4.2 Prepad
A prepad is not necessary for the calibration test.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 19 of 81

2.4.3 Closure Pressure less than Hydrostatic Pressure


Fluid will flow from the wellbore into the fracture during closure if closure pressure is
less than hydrostatic pressure. Calculate the quantity of fluid displaced when
closure pressure occurs. If the displaced fluid is more than 10% of the fracture
volume at shut-in (volume injected times efficiency), a special wellbore isolation tool
should be used in conjunction with a wireline-conveyed bottomhole pressure gauge.
Such tools have been used before but may have to be specially constructed. A
bottomhole pressure gauge must always be used in these cases.

2.4.4 Post-Job Wireline Surveys


Postjob logs should not be run until closure has occurred and pressure monitoring
has ceased. Cable movement in the wellbore and fluid drag on the cable can affect
the pressure decline data. If postjob logs are to be run, consider using a wireline
conveyed bottomhole pressure gauge set below the perforations.

2.5 Terminology

2.5.1 Fracture Extension Pressure


The fracture extension pressure is the pressure required to extend an existing
fracture. Typically, the fracture extension pressure is 50 to 200 psi greater than the
closure pressure because of fluid friction in the fracture and fracture toughness.

2.5.2 Initial Shut-in Pressure


The initial shut-in pressure provides an upper bound for the determination of closure
pressure.

2.5.3 Closure Pressure


An accurate determination of the closure pressure is essential for an analysis of the
fracturing pressure because it is the datum for determining the net pressure. The
closure pressure is the fluid pressure at which the fracture closes (zero width). This
pressure is equal to, and counteracts, the minimum principal stress in the rock that is
perpendicular to the fracture plane. The closure pressure reflects a global average
of the minimum stress, which is a local quantity and is not constant over the zone of
interest. The closure pressure generally is less than the breakdown pressure
required to initiate a fracture and always less than the fracture extension pressure.

2.5.4 Rebound Pressure


The rebound pressure after shut-in is a lower bound of the closure pressure.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 20 of 81

2.6 Equipment Requirements

2.6.1 Monitoring Equipment


An on-site MicroVAX1 computer is not absolutely necessary for performance of the
pumping portion of the DataFRAC service. However, a MicroVAX will greatly
enhance data manipulation and examination. A MicroVAX computer is necessary if
onsite data analysis and treatment design using the FracCADE software is desired.
There are two alternatives if a MicroVAX computer is not available.
1. Perform a hand analysis.
2. Perform the analysis in the office. This option may force a redesigned treatment
to be pumped at some later date.
Two French curves are helpful for determining the reversal in curvature (closure
pressure) from the flowback pressure plots. Always carry linear graph paper for any
hand plotting needed as well as log-log paper for plotting a Nolte-Smith plot if
necessary.

2.6.2 Pumping Equipment


Diesel-powered pumpers are recommended for the closure test. Turbine powered
pumpers are not recommended for the closure test because rate control is poor,
especially at low pressures. Any type of pumpers may be used for the calibration
test.

2.6.3 Pressure Measuring Equipment


Accurate pressure measurement is critical to the success of the DataFRAC Service.

2.6.3.1 Surface Measurement Methods


For the pressure-decline phase of the calibration test, the bottomhole pressure can
be calculated from the surface pressure as long as the fluid density is constant and
the bottomhole pressure is greater than the hydrostatic pressure. The main problem
with using the treating pressure for analysis is that the friction pressure makes the
Nolte-Smith plot less accurate and can indicate erroneous trends. In the overall
analysis, the Nolte-Smith plot is very valuable if accurate bottomhole pressure and
closure pressure are used.
A good method for measuring bottomhole pressure is with a “live” annulus or a
“dead-string tubing” and a homogeneous fluid. This eliminates friction pressure
calculations. With a known hydrostatic pressure, bottomhole pressure can be
accurately calculated. The density of the static column of fluid must be known
(circulate the well and check the specific gravity of the fluid prior to injection). The
fluid must not contain any trapped gas. This method is generally adequate for wells

1 Trademark of Digital Equipment Corporation

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 21 of 81

with a bottomhole static temperature less than 250°F (121°C) and a depth less than
10,000 ft.
Significant hydrostatic pressure changes may result from a change in fluid density
during closure in deep, hot wells. This occurs when the wellbore fluid is warmed by
the formation. After pumping, surface pressures can actually increase while the
bottomhole pressure decreases (Fig. 8). In a 16,000 ft, 325°F (163°C) well,
hydrostatic pressure change can be as much as 250 psi for water (Fig. 9). The
effects on oil will be much greater because of the greater thermal expansion of oil.
This compromises any results from surface readings because overly optimistic fluid-
loss and efficiency values will be implied. Therefore, the use of surface readings for
deep, hot wells is not acceptable.

Fig. 8. The change in surface pressure during closure in deep, hot wells.
THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 22 of 81

Fig. 9. Hydrostatic head changes during closure.


If closure pressure is less than the hydrostatic pressure of the injected fluid, then
pressure analysis is not possible from surface measurement and a wireline-
conveyed bottomhole pressure gauge must be used.

2.6.3.2 Bottomhole Pressure Gauge Measurement


The best choice for measuring bottomhole pressure is with a bottomhole pressure
gauge thereby eliminating friction calculations and hydrostatic considerations. For
fluids without proppant, this can safely be done with a wireline-conveyed gauge, in
the fluid stream if necessary. To ensure the wireline tension does not exceed a safe
level, the increased tension due to fluid drag must be calculated using Eq. 1 before
the job begins.

π
T= × d ID × d w × Pf (1)
4
Where:
T = tension due to fluid drag (lbf)
dID = inside diameter of pipe (in.)
dw = diameter of wire (in.)
pf = estimated total friction pressure in pipe (psi).
Wireline tension must be calculated and confirmed to be safe with the wireline
service company prior to rig-up to avoid parting the wire and subsequent job failure.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 23 of 81

A wireline-conveyed bottomhole pressure gauge interfaces through a Remote Data


Acquisition (RDA) box. Voltage and frequency inputs for the RDA box are:
• 0-20 mA
• 4-20 mA
• 0-4000 Hz (12 volts)
Additional information is provided in the PPR System Operator's Manual.
Use of the PPR* pumping parameter recorder or other monitoring device is
suggested. The wireline-conveyed pressure gauge should be tested prior to job
execution.

2.6.4 Treating Equipment


Wellhead rig-up requirements must be considered and communicated to the wireline
service company. If the injection rate through two-inch treating equipment is greater
than 8.5 bbl/min, a frac cross may be necessary. At rates less than 8.5 bbl/min, a
lateral may be sufficient. The Dowell Location Safety Standards manual provides
the maximum pumping rates through treating equipment.

2.6.5 Flowback Equipment


Flowback rate must be monitored accurately for adequate control. Response time
on the flowmeter should be 3 sec or less.

2.6.5.1 Magnetic Flowmeters


Magnetic flowmeters are used in conjunction with water-base (conductive) fluids.
The Dowell Flumag flowmeter is commonly used. Other magnetic flowmeters may
be used. Magnetic flowmeter information is provided in the Sensors Verification
Guide.

2.6.5.2 Turbine Flowmeters


Turbine flowmeters are typically used with oil-base (nonconductive) fluids, but may
be used with any fluid type. Turbine flowmeter information is provided in the
Sensors Verification Guide.

2.6.5.3 Chokes and Gate Valves


An adjustable choke or a gate valve is commonly used to regulate flowback rate.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 24 of 81

3 Execution
Treatment design for the closure test and the calibration test is provided in Section 2.
Location Safety Standard Number 5, 5A, and 5B provides procedures for approved
on-location practices.

3.1 Pre-Performance Guidelines


Certain guidelines are common for both the closure test and the calibration test.
1. Equipment is rigged-up in accordance with Location Safety Standard Number 5,
5A, and 5B. Additional details for equipment rig-up and flowback are provided in
Fig. 10 (conductive fluids) and Fig. 11 (nonconductive fluids). An adjustable
choke or a gate valve is used in place of the choke nipple in the flowline
(bleedline).
2. If a static string is used, ensure the static fluid column is filled with a fluid of
known specific gravity with no gas cap. The preferred method is to circulate
from the tubing to the annulus at high velocity.
3. Ensure that suction hoses, discharge hoses, manifolds, pumps, blenders, and
discharge piping do not contain proppant.
4. Backup pressure transducers must be rigged-up and calibrated. Do not provide
any more than one display for the same pressure. The transducers are accurate
to 1% of full scale. This means a 15,000 psi transducer is accurate to ±150 psi.
If the maximum pressure will be low, suggest using a 0 to 5,000 or 0 to 10,000
psi transducer for better accuracy. Do not allow anybody to hammer on
transducers during any phase of testing.
5. The recording period for data acquisition should be 5 to 15 sec. High
permeability formations and/or low-volume (short closure time) pump tests
require a shorter time interval (5 sec or less). Do not set a PPR to record data
from the POD* blender or the storage capacity of the tapes will be exceeded.
During the pressure decline, do not allow pausing or constant changing of
calculated data.
6. Determine the expected closure pressure. The closure pressure may be
approximated using Eq. 2.
Approximate closure pressure =
Overburden pressure + ( Reservoir Pr essure × 2
3 (2)

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 25 of 81

7. If the wellbore is full of fluid, note the initial bottomhole pressure. Otherwise, note
the quantity of fluid required to fill the wellbore (pressure rise). Once the wellbore
is full, shut down and record the pisi. Calculate bottomhole pressure using the
initial fluid level.

Fig. 10. The DataFRAC Service rig-up when pumping conductive fluids.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 26 of 81

Fig. 11. The DataFRAC Service rig-up when pumping nonconductive fluids.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 27 of 81

3.2 Closure Test


The general steps in a closure test are
1. Rig-up, mix fluid, and perform quality control activities.
2. Perform the step-rate phase of the closure test.
3. Perform the flowback phase of the closure test.
4. Perform a modified step-rate phase if necessary.
5. Perform a modified flowback phase if necessary.

3.2.1 Step-Rate Phase


Step-rate phase guidelines are:
1. The pump operator should know the pump gear and speed for each of the steps
prior to pumping operations. This will facilitate rapid step-rate changes. Getting
the injection rate (as well as flowback rate) established quickly must be stressed.
Exact rates are not important  constant rates are. Fluid-end ratings and
constants are provided in the Treating Equipment Manual. Pump performance
curves are provided in the appropriate pumping equipment operators manuals.
2. Take pressure readings after establishing a new pump rate (prior to increasing
the rate again).
3. Determine if fracture extension is occurring during the last injection stage by
plotting rate versus pressure. This will indicate fluid loss to the matrix leakoff or
fracture extension (Fig. 2). Fracture extension pressure will be 50 to 200 psi
greater than the closure pressure. Remember to plot rate versus bottomhole
pressure (not treating pressure). If treating pressure is plotted, the friction
pressure will distort the values at higher rates and produce erroneous results.
4. Increase the pump rate during the last stage if fracture extension is not
occurring. If fracture extension is occurring, terminate the stage after the
desired length of time. Water hammer effects can be minimized by reducing the
pump rate to 10% of the final rate for 10 to 15 sec before shutdown.
5. Determine the true perforation friction pressure using Eq. 3 and Fig. 12, Fig. 13,
Fig. 14, Fig. 15, Fig. 16, or Fig. 17. Using Fig. 18, determine the estimated
perforation friction pressure if all perforations were open. If the true perforation
friction pressure is greater than twice the estimated perforation friction pressure,
wellbore restriction is too great and should be reduced by pumping a diverting
treatment or reperforating. Injecting small quantities of proppant near the end of
the pad of the proposed fracturing treatment may erode the restriction.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 28 of 81

p pf = pw − ptf − pisi (3)

Where:
ppf = perforation friction pressure (psi)
pw = surface fracturing pressure (psi)
ptf = tubular friction pressure (psi)
pisi = initial shut-in pressure (psi).

Fig. 12. Friction pressure of water in the tubing and casing.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 29 of 81

Fig. 13. Friction pressure of water in the annulus.

Fig. 14. Friction pressure of brine in the tubing and casing.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 30 of 81

Fig. 15. Friction pressure of brine in the annulus.

Fig. 16. Friction pressure of diesel in the tubing and casing.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 31 of 81

Fig. 17. Friction pressure of diesel in the annulus.

Fig. 18. Flow rate versus differential pressure in perforations.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 32 of 81

3.2.2 Flowback Phase


Flowback must be initiated at a constant rate as soon as possible. Remember to
isolate the pump(s) from the well. Do not allow bottomhole pressure to fall below
200 psi above the initial bottomhole pressure. Do not flow-back more fluid than was
pumped.
Note the indicated change in bottomhole pressure during shutdown and calculate the
perforation friction pressure. If the perforation friction pressure is more than twice
the expected amount, discuss the discrepancy with the client.

3.2.2.1 Flow Control


Adjustment of the choke or valve may be accomplished using one of two methods.
1. Pump through the choke or valve prior to performing the step-rate/flowback test
to preset the choke or valve. The choke or valve is adjusted to the desired rate
when flowback is initiated.
2. Adjust the choke or valve during the last pumping stage of the step-rate test.
The pump rate through the choke or valve will be in addition to the pump rate
required for the last stage.
Flowback rate accuracy is not critical; ± 20% is acceptable. However, a constant
flowback rate is critical.
Table 1 provides approximate choke settings (using a 15,000 lbf adjustable choke,
part number 515077000) for flowback of oil-base fluids. Table 2 provides
approximate choke settings for flowback of water-base fluids. Verify the setting by
pumping through the choke at the anticipated flowback rate and pressure shut-in
pressure. This is a good time to functionally check the flowmeter.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 33 of 81

Table 1. Approximate Choke Settings For Flowback Of Oil-Base Fluids (Sg = 0.7)

Pressure Flow Rate (bbl/min)


(psi)
1 3 15 10 15 20
1,000 14 24 31 44 53 62
2,000 12 20 26 37 45 52
3,000 10 18 23 33 41 47
4,000 10 17 22 31 38 44
5,000 9 16 21 29 36 41
6,000 9 15 20 28 34 39
7,000 8 15 19 27 33 38
8,000 8 14 18 26 32 37
9,000 8 14 18 25 31 36
10,000 8 13 17 25 30 35
11,000 8 13 17 24 29 34
12,000 7 13 17 23 29 33
13,000 7 13 16 23 28 33
14,000 7 12 16 23 28 32
15,000 7 12 16 22 27 31

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 34 of 81

Table 2. Approximate Choke Settings for Flowback of Water-Base Fluids (Sg = 1.0)
Pressure Flow Rate (bbl/min)
(psi)
1 3 15 10 15 20
1,000 15 26 34 48 58 68
2,000 13 22 28 40 49 57
3,000 11 20 26 36 44 51
4,000 11 18 24 34 41 48
5,000 10 17 23 32 39 45
6,000 10 17 22 31 37 43
7,000 9 16 21 29 36 42
8,000 9 16 20 28 35 40
9,000 9 15 19 28 34 39
10,000 8 15 19 27 33 38
11,000 8 14 19 26 32 37
12,000 8 14 18 26 31 36
13,000 8 14 18 25 31 36
14,000 8 14 17 25 30 35
15,000 8 13 17 24 30 34
The downstream 1 x 2 hamer valve (control valve) in the flowline (bleedline) may be
used for flow control if the adjustable choke becomes plugged and can not be
cleared. Use the hamer valve for flow control only as a last resort.
The choke (or valve) operator must have a rate display for reference. Relaying rates
via radio is not acceptable.

3.2.2.2 Flowmeters
When using a turbine flowmeter, open the control valve slowly to avoid a fluid surge
and subsequent flowmeter damage. Never allow a low-pressure magnetic flowmeter
(for example, Fischer-Porter) to be placed upstream of the choke. Flowmeters must
have a full pipe of flow to maintain accuracy. A backup flowmeter is recommended.

3.2.3 Closure Test Modifications


Modifications to the closure test may be required for the following reasons.
• Extension pressure was not attained.
• An overshoot of fracture extension pressure took place.
• Flowback rate was inaccurate.
Section 2.2 provides design modifications for the closure test.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 35 of 81

3.3 Calibration Test


The general steps in a calibration test are
1. Rig-up, mix fluid, and perform quality control activities.
2. Perform the injection phase of the calibration test.
3. Perform the pressure-decline phase of the calibration test.

3.3.1 Injection Phase


The type of fluid and injection rate for the calibration test are the same as the type of
fluid and injection rate of the proposed fracture treatment.
Injection phase guidelines are
1. If the flush fluid volume is more than 10% of the calibration fluid volume, the
treatment fluid should be circulated to the top perforation. If circulation is not
possible, pump the tubing volume (or annular volume, whichever is applicable)
at low rate. Stop pumping and let the pressure fall below closure before
resuming pumping. Fluid warming will change the fluid characteristics. Do not
wait any longer than necessary if the well has a high bottomhole static
temperature.
2. When using crosslinked fluids, accurate crosslinker/activator additive rate is
especially critical for correct DataFRAC analysis. A linear fluid, as opposed to a
crosslinked fluid, will cause a different pressure response and have different
fluid-loss characteristics. A back-up additive pump is recommended.
3. Use the closure pressure determined from the closure test in calculation of net
pressure for the Nolte-Smith plot. Reset pump time to zero when fluid enters the
perforations and start the plot.
4. Calculate fluid friction pressure using bottomhole pressure or obtain the shut-in
pressure during the calibration test. Initial shut-in pressure obtained after
pumping the flush fluid yields friction pressure for the flush fluid, not the
calibration fluid.
5. Stop pumping when flush is complete. Reduce water hammer effects by
reducing the pump rate to 10% of the final rate for 10 to 15 sec before shutdown.
6. Record the shut-in pressure when the pump rate falls to less than 2% of the
treatment pump rate.
7. Isolate the pumping equipment when all pumping has stopped.
8. Calibration tests using foamed fluids must be flushed with a linear fluid not
containing carbon dioxide, nitrogen, or crosslinker/activator.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 36 of 81

3.3.2 Pressure-Decline Phase


Pressure-decline phase guidelines are
1. Monitor pressure decline for 1.25 times the closure time or for twice the injection
time, whichever is longer. Recording closure is very important.
2. Do not allow anybody to hammer on the line or disturb the transducers during
monitoring activities.
3. Do not run postcalibration-test wireline surveys during monitoring activities.
4. If the annulus is isolated, do not reduce or increase pressure during monitoring
activities. Expansion or contraction will affect the tubing pressure and the final
analysis if surface pressure is used.

3.3.3 Contingency Plans


1. If an operational problem occurs with less than 30% of the fluid volume pumped,
stop pumping and correct the problem. Resume pumping the remaining fluid at
the design rate. Do not continue pumping at a reduced rate. Do not be
concerned about a fluid leak unless the leak causes safety concerns or is
tremendous, (gallons/minute). The volume loss compared to the leakoff in the
fracture is small and will not affect the pressure decline.
2. If an operational problem occurs with approximately 50% of the fluid volume
pumped and the problem can be corrected quickly, stop pumping and note the
loss of net pressure. If more than 20% of the net pressure is lost, consider
a) starting over
b) monitoring the pressure decline and pumping a second calibration test with
the remaining fluid.
If less than 20% of the net pressure is lost, resume pumping and analyze using
the total volume pumped and the final injection rate. The pump time will be filled
in on the DataFRAC form.
3. If an operational problem occurs with more than 70% of the fluid volume pumped,
stop pumping and monitor the pressure decline. Be sure to use the actual
volume of fluid injected into the formation in the analysis.
At least 50% of the total volume should be pumped at the designed rate.

4 Evaluation
The DataFRAC analysis consists of three essential parts.
1. closure test for closure pressure
2. calibration injection for fracture geometry
3. calibration decline for fluid-loss behavior

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 37 of 81

For correct analysis, the actual bottomhole pressure (BHP) must be used (See
Ref. 2: Chapter 7.6.2). Combining the analysis of the closure test, pressure during
pumping (as predicted by a fracture simulator) and pressure decline during closure
provide a consistent interpretation and the enhancement of the three parts.
Consistent values of the fracturing parameters for all the three analysis provide a
sound basis for proper DataFRAC evaluation and subsequent treatment design.

4.1 Closure Test Analysis


The closure pressure is the fluid pressure for which the fracture effectively closes
without proppant. The closure pressure is distinguished from the minimum stress.
The stress is a local parameter which can vary over the pay zone, whereas the
closure pressure is a global parameter reflecting the gross behavior of the pay zone.
The field procedures for the closure pressure test require the creation of a fracture in
the complete zone as opposed to a “micro” fracture for the stress test.
The methods used for determining the closure pressure include the step rate and
flowback test.
The step rate is analyzed using a BHP versus rate plot and the flowback is analyzed
using a BHP versus time plot.

4.1.1 Step Rate  The BHP-Versus-Rate Plot


The BHP-versus-rate plot (Fig. 2 and Fig. 3) should show two different slopes
indicating matrix leakoff at low pressures/rates, and fracture response at higher
pressures/rates. The extension pressure provides an upper bound for the closure
pressure and defines the required range of pump rates for extending a fracture in the
zone.

4.1.2 Flowback  The BHP-Versus-Time Plot


The inflection point from concave up to concave down on the BHP-versus-time plot
(Fig. 19) of the flowback response, is the point of increased pressure drop through
the entrance of the fracture. The lowest point of the pressure derivative curve will be
the inflection point. Several publications prior to 1993, indicated closure occurred at
the inflection point. Subsequent analysis, with a comprehensive fracture simulator,
indicated closure pressure occurs at a lower pressure and near the intersection of
the tangents shown in Fig. 19.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 38 of 81

Fig. 19. Flowback test (after Nolte, 1982/1994).

4.1.3 Confirmation of Closure Pressure


The confirmation of closure pressure can be done using the square-root of time plot
or G-plot during the shut-in of the calibration treatment. The closure pressure is
inferred as change of the slope on either of these plots (Fig. 20). This method
normally does not provide a definitive indication of the closure pressure because of
the existence of multiple slope changes. The fracture closure generally causes one
of the slope changes in the BHP versus: t plot. A change in slope of the “G” plot
also is a typical indication of closure pressure.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 39 of 81

Fig. 20. Effect of closure on BHP versus square root of t and G- plots.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 40 of 81

4.1.4 Rebound Pressure


After the pressure drops below the estimated closure point during flowback, the well
is shut-in and the rebound pressure is monitored. The rebound pressure provides a
lower bound of the closure pressure and the inflection point provides an upper bound
of the closure pressure (Fig. 21).

Fig. 21. Rebound pressure; lower bound of closure pressure.

4.2 Calibration Injection for Fracture Geometry


The Nolte-Smith plot (log-log plot of the net pressure versus pumping time) provides
an important diagnostic tool for determining how the fracture is propagating and the
fracture geometry during pumping. The analysis enables the simulation and
calibration of the pressure with a numerical fracture simulator and permits
reconciliation of the ideal assumptions and actual field conditions. The magnitude of
the net pressure from the fracture simulator permits a verification of fracture
parameters such as modulus, height, toughness or barrier stress difference.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 41 of 81

4.2.1 Elastic Fracture Compliance


For fracturing applications, the linear elastic assumption of Sneddon's classical
solution is applied. From the solution, the average fracture width can be expressed
in terms of the closure pressure (pc) fracture compliance (cf) and net wellbore
pressure (∆pf) as:
w = c f ∆p f = c f ( p f − pc ),
Where:
πβd
cf =
2 E'

Fig. 22 indicates that the behavior of a pressurized crack is analogous to a pre-


loaded spring.
pc = “spring pre-loaded”

1 E'
∝ = “spring constant”
cf d

___
β = ∆p f
Pw − Pc (see Section 4.3.8.4)
E = rock modulus
d = “characteristic” dimension of frac geometry

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 42 of 81

Fig. 22. Analogy of a pressurized crack to a pre-loaded spring.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 43 of 81

The average crack width can be expressed by Sneddon's relationship in terms of “d”,

πβd ( p f − pc )
w=
2 E'
This relationship is used to model a fracture as follows:
• PKN: xf
→ ∞ d = hf
hf
hf
• KGD: → ∞ d = 2x f
xf
2x f 32
• Radial: →1 d= R ≈ R and x f = R.
hf 3π 2
The KGD model is more appropriate when the fracture length is smaller than the
height, while the PKN model is more appropriate when the fracture length is much
larger than the height. The radial model is most appropriate when 2xf is about equal
to the height.

4.2.2 Pressure During Pumping

4.2.2.1 Fluid Flow and Pressure in Fracture


The pressure gradient in the fracture can be expressed as;
n′
dp K'  q 
∝ 2n' +1  i  .
dx w  hf 
This expression relates the gradient down the fracture length to the fluid velocity or
flow rate. Introducing the fracture compliance (w = cf∆ pf), integrating along the
fracture length and assuming ∆pf = 0 at the tip, results in;
1
 K '  q  n'  ( 2n' + 2)
∆p f ∝  2n' + 1  i  x f 
c f  hf  
1.  
1
 n'  2n
 qi 
w = c f ∆p f ∝  c f K ' x f   
  hf  

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 44 of 81

These proportionalities indicate the effect on pressure and width from variations of
fluid rheology, injection rate, fracture geometry and models (in terms of fracture
compliance). Substituting the appropriate compliance relationship for the three basic
models gives;
1
• PKN:  x f  ( 2n' + 2)
∆p f ∝ A 3n' +1  ,
 h f 
1

• KGD:  1  ( 2n' + 2)
∆p f ∝ A n 2n'  ,
 h f x f 
1
 1  ( 2n' +2)
• Radial: ∆p f ∝ A 3n'  .
R 

1
2n' +1 n ′ ( 2n' + 2)
Where: A = (E K ' qi ) ,
which is the same for all the three models. The relationships also indicate that with
increasing penetration, the net pressure increases for PKN model and decreases for
the KGD and radial models.
For constant injection rate, the fracture growth can be expressed in terms of time
and bounded by two extreme cases for fracture efficiency, η:
• Upper bound: No fluid loss (that is, Vf = Vi = qit). V f = w A f ∝ t ; η → 1

• Lower bound: Almost total fluid loss (that is, VL → Vi = qit and Vf → 0).

A f ∝ t 1/ 2 ; η→0
A f = fracture face area.

The fracture penetration increases with time and depends on the fluid loss during
injection. By combining the bounds for time dependence of penetration, the
relationship for net pressure and width, the net pressure yields;
• PKN: ∆p f ∝ t 1/ 4( n' +1) ( η → 0)
∆p f ∝ t 1/( 2n+ 3) ( η → 1)

∆p f ∝ t − n'/2 ( n' +1) ( η → 0)


• KGD: ∆p f ∝ t − n'/( n' +2) ( η → 1)

∆p f ∝ t −3n'/8 ( n' +1)) ( η → 0)


• Radial: ∆p f ∝ t − n'/( n' +2) (η → 0)

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 45 of 81

The previous expressions for pressure assume the fluid viscosity dominates the
pressure distribution and ignores the fracture toughness of the formation. This
assumption is generally valid for fractures with dimension in excess of 50 ft using
high-viscosity fluids. For the case of small-scale fractures created with low viscosity
fluids, fracture toughness can dominate and result in different exponents for time.
The expressions for the net pressure are all exponential expressions. As a result, a
log-log plot of net pressure versus time should yield a straight line with slope equals
to the respective exponents: positive for PKN and negative for KGD and radial
models. The log-log plot of net pressure versus time as introduced by Nolte and
Smith, forms a basis for the interpretation of pressure data during fracturing.

4.2.2.2 Nolte-Smith Plot and Evolution of Pressure During Pumping


Fig. 23 shows the evolution of the fracture geometry and the Nolte-Smith plot for an
ideal case with bounding formations of higher stress. During the initial phase of
propagation (stage 1), the fracture area increases in the radial mode (point source)
or as expanding ellipses (line source). The line source can be approximated by KGD
model. For this initial phase, the log-log slope is negative and between -1/8 and -
1/4. This phase continues until the fracture is affected by barriers, which may occur
after a very short time.

Fig. 23. Evolution of fracture geometry and pressure during pumping.


The fracture will then propagate in PKN mode after the radial model encounters
barriers above and below (stage 2) which results in increasing pressure and the log-
log slope is between 1/4 and 1/8. Without proppant, the net pressure is limited to a

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 46 of 81

value slightly below the stress difference (∆σ) of the barrier being penetrated. At this
time, the height begins to increase significantly and the pressure would be
approximately constant (stage 3).
Nearly constant pressure indicates the pressure capacity for the formation, which is
determined by in-situ stress difference. When the net pressure reaches this
capacity, fracture extension becomes relatively inefficient, as discussed in the
following sections.

4.2.3 Deviations from Ideal Geometry

4.2.3.1 Height Growth


Height growth into stress barriers is a common deviation from the ideal PKN model.
Fig. 24 illustrates the pressure and vertical cross section of the width profile. Stage
“a” is the PKN propagation stage. The positive log-log slope will continue until the
net pressure approaches the stress difference of the barrier. At this stage (stage
“b”), the height will increase and the pressure would be approximately constant.
During stage “c”, the barrier is crossed and the fracture enters a lower stress zone
resulting in an accelerated rate of growth at decreasing pressure and width in the
primary zone. The width profile indicates that a “pinch point” occurs in the barrier
after stage “b”. The pinch point has essentially no width during the transition from
stage “b” to stage “c”. The pinch point can cause proppant to bridge as fluid is
permitted to pass through. The resulting excessive dehydration of the slurry coupled
with the decreasing width can result in a rapid screenout even at low proppant
concentration.

Fig. 24. Pressure and width for height growth through barriers
(after Nolte, 1989).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 47 of 81

The slurry dehydration, decreasing width, and height growth can be reduced by the
following methods:
1. Place an impermeable mixture of proppant between the pad and the proppant
stages to form an impermeable bridge at the pinch point.
2. Pumping a pre-treatment with a diverting agent (INVERTAFRAC or
DIVERTAFRAC).

4.2.3.2 Fissures
Another possible cause for a period of constant pressure is the opening and inflating
of natural fissures. Pressure-dependent leakoff due to fluid loss into fissures is
thought to contribute to screenouts in low permeability formations where limited fluid
loss would otherwise be anticipated.
Two fissure models have been reported.
1. Slight permeability enhancement
The permeability enhancement is not significant until the effective stress
becomes negative and the fissure aperture opens. At this time, fluid loss
becomes significant and regulates the pressure to a constant value.
2. Highly stress-sensitive permeability and fluid loss
The permeability and fluid-loss enhancement are significant throughout the
treatment, with the effect accelerating as the pressure increases. If the
treatment continues, the negative effective stress condition can occur with
constant pressure.

Fig. 25. Pressure and width for opening natural fissures


(after Nolte, 1989).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 48 of 81

Fig. 25 illustrates the pressure response and the horizontal cross section of the width
profile. The secondary fracturing occurs in natural fissures or cracks which are
crossed by the primary fracture. These feature normally have relatively higher
permeability than the matrix and the fluid leakoff is high.
The fissures will open when the fluid pressure exceeds the formation stress acting
across them.
σ − σ1
∆p f > 2 ≈ 1.5 ∆σ H
1 − 2ν

∆σH = σ2 - σ1 = horizontal stress difference.


This implies that effective fracturing will require a significant stress difference
between the principal horizontal stress to avoid opening of natural fissures. When
this magnitude of pressure is reached, the fissures open and act to regulate the
constant pressure at this critical magnitude. A significant portion of the injected fluid
can be lost because of a large number of fissures that can open at this critical
pressure. The accelerated fluid loss can lead to excessive slurry dehydration and a
screenout (stage “c” of Fig. 25).
The accelerated fluid loss can be reduced using the following methods.
1. Before the fissure aperture opens, use very fine particles (for example,
300-mesh particles) in the pad.
2. After the fissures open, and maintain constant pressure, use 100-mesh particles
between the pad and proppant stages (Note: 100-mesh particles can screenout
the treatment when they reach the tip).

4.2.3.3 T-Shape Fracture

Fig. 26. Pressure and width for T-shape fracture


(after Nolte, 1989).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 49 of 81

When the fracturing pressure is greater than overburden stress, a fracture can
propagate in both the horizontal and vertical planes. This geometry is called a
T-shape fracture and the pressure response and a vertical cross section of the width
profile are illustrated in Fig. 26. The figure indicates stage “c” has a near constant
pressure response. The horizontal component growth requires pressures greater
than the overburden pressure and occurs at;

∆p f ≥ OB − pc

Where:
OB = vertical overburden stress
pc = closure pressure.
The width of the horizontal fracture component will be narrow and have twin pinch
points at the juncture with the vertical component. The limited width of the horizontal
component can restrict proppant entry, excessively dehydrate the slurry in the
vertical component, and lead to premature screenout.
The T-shape fracture is the easiest to diagnose: Bottomhole injection pressure
approximately constant at a value slightly above the overburden pressure (that is,
about one psi/ft of true vertical depth).

4.2.4 Pressure Capacity


Summarizing the prior sections (using Fig. 27) a period of constant pressure for a
vertical fracture can occur because 
• The pressure approaches the stress of a barrier and causes significant height
growth;
∆p f ≤ ∆σ ν

∆σv = barrier stress difference.


• The pressure exceeds the stress acting on natural fissures and the fissures open;
∆σ H
∆p f ≈ .
1 − 2ν

• The pressure exceeds the overburden pressure, and the initiation of T-shape
fracture begins;
∆p f ≥ OB − pc

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 50 of 81

Fig. 27. Definition of pressure capacity from in-situ stresses.


For these cases, the limiting pressure is called the formation pressure capacity. The
formation acts as a pressure vessel with a pressure capacity defined by stress
differences. Exceeding the pressure capacity leads to inefficient extension due to
height growth, the formation of T-shape fracture or fissures opening.

4.2.5 Near-Wellbore Restriction


High near-wellbore pressure losses sometimes experienced during the hydraulic
fracturing treatment should be considered in fracturing pressure analysis, that is,
subtracted for determining net pressure. In addition to inadequate perforating, a
potential cause of high near-wellbore pressure losses is that the well and the fracture
plane are not aligned, that is, on deviated wells or wells close to faults (deviated
principal stress). For these cases, the fracture initially aligns with the wellbore, and
then turns to align normal to the far-field minimum stress. The fracture entrance
experiences a normal stress greater than the minimum stress, leading to a fracture
width restriction and increased pressure losses within the entrance.
The stress state within the entrance is illustrated by Fig. 28 and the Mohr circle in
Fig. 29.
AB = fracture plane
σ1 = minimum principal stress
σy = stress parallel to the wellbore
σx = stress normal to the wellbore.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 51 of 81

Fig. 28. Stress state within the entrance of deviated well or stress.
For the deviated stress case, σy and σx are equal to overburden and horizontal
stress, respectively. The principal (that is, minimum and maximum) stresses are not
horizontal or vertical and the fracture is inclined. For the deviated well case, the
principal stresses are assumed horizontal and vertical, σy and σx are parallel and
normal to the inclined wellbore. σx can be estimated as the sum of the minimum
stress (that is, closure pressure) and the apparent near-wellbore friction, pwf; that is,

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 52 of 81

σx = σ1 + pwf with σ1 estimated from a closure pressure test and pwf from the BHP
change during a shut-in.

Fig. 29. Mohr circle of deviated well or stress.


Using the Mohr's circle in Fig. 29, information about the state of stress and relative
fracture orientation can be obtained from known information (for example, θ for
deviated well, or σy = OB for vertical well).

Radius of Mohr circle, σx + σy


R= − σ1
2

Therefore;
R − pwf
cos 2θ = ,
R
Where:
θ = angle between wellbore axis and fracture plane
pwf = near-wellbore friction pressure.
Significant entrance friction can be diagnosed by a large difference in the bottomhole
injection pressure during fracturing and the initial shut-in pressure (ISIP). The
entrance friction responds the same as perforation friction and tends to decrease
when proppant is added. For pre-fracture tests, a significant entrance restriction can
be indicated by a large difference (for example, greater than 200 psi) between the
extension pressure (Fig. 2 and Fig. 3) and closure pressure (Fig. 20 and Fig. 21).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 53 of 81

4.2.6 Fracturing Pressure Interpretation Summary

Fig. 30. Nolte-Smith plot of fracturing pressure.


Small Positive Slope
The fracture is propagating under contained height and free lateral extension in a
manner similar to PKN model. The approximate log-log slope is 1/8 to 1/4.
Zero Slope
Reduced penetration rate potentially caused by: height growth, fissures opening or
the formation of T- shape fracture. The constant pressure during this period is called
formation pressure capacity which is determined by the in-situ stresses; and hence
likely to be the same for offset wells.
Positive Slope
Flow restriction causing fracture width to be increased with limited extension,
potential proppant bridging and screenout.
1− η
• Tip screenout condition, the log-log slope ≈ 1 + 0.64 η ∆t D

where η is the efficiency at screenout and ∆tD is the time after screenout divided
by the screenout time.
• Log-log slope > 1 indicates restriction in the fracture.
• Log-log slope >>1 (very high slope) indicates restriction near or at the wellbore
resulting from a near-wellbore restriction (Section 4.2.5) or exceeding the
pressure capacity (Section 4.2.4).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 54 of 81

Negative Slope
Unrestricted height growth through a lower stress formation. It also indicates growth
of a radial fracture with the fracture propagating in a manner similar to the KGD or
radial model.

4.2.6.1 Example of Radial Fracture


Fig. 31 shows the interpretive log-log plot of a radially propagating fracture from a
calibration treatment of a massive chalk section in the North Sea. The plot shows
the expected behavior of a radial fracture with a log-log slope equal to about – 1/8.
The period of lower pressure at about 10 min resulted from a shut-in. The small
pressure change of about 50 psi indicates there was no significant entrance
restrictions.

Fig. 31. Net pressure with radial fracture (after Smith et al. 1987).

4.2.6.2 Simulation of Pressure During Pumping and Decline


The numerical simulation of the pressure response is an important tool in the
calibration injection for fracture geometry and calibration decline, for cases in which
the idealized 2D geometry models are inadequate. These nonideal cases may
include height growth, stress sensitive fissures, and fracture penetration and
recession during the decline.
The log-log plot of the fracturing pressure is generally a qualitative and diagnostic
tool. Quantitative interpretation can be obtained by comparing the predictions from a

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 55 of 81

numerical fracture simulator and the actual treatment. The net pressure is primarily
governed by the rock mechanical properties and is relatively insensitive to rate and
viscosity. Numerical simulation is used to calibrate or confirm values for the
parameters that govern the pressure response, that is, fracture compliance, rock
stress difference and fluid-loss coefficient. The calibrated parameters can then be
used to make design changes in subsequent treatments. Even though there is not a
unique set of fracturing parameters that match a pressure response, a calibrated set
provides a rational basis for more effective treatment design.

Fig. 32. Measured and simulated net pressure: opening natural fissures
(after Nolte, 1982).
Fig. 32 shows an example of the reported application of fracturing pressure
simulation during injection and decline. The pressure plot shows a near-constant
pressure period indicating that the formation capacity has been reached. The
pressure capacity value of 1700 psi, because of the opening of natural fissures,
lasted for about 100 minutes prior to shut-in. The pressure capacity is governed by
the rock stress and should be expected to be similar throughout the field, provided
there are no significant lithological or tectonic changes. Consequently, once the
pressure capacity is determined and the pressure calibrated by the simulator,
rational design changes can be made for more effective treatment for the remainder
of the wells in the field, that is, design using a pressure calibrated simulator to stay
below the pressure capacity for more efficient penetration. In addition, the pressure
simulation during decline may provide consistent interpretation and result in
enhancement of both injection and closure analysis; however, an appropriate
numerical fracture simulator is required to correctly apply these concepts. Ideally the
simulator should include the effects of spurt loss, pressure-dependent fluid loss (with
and without sensitive fissures), fluid temperature and compressibility, poroelastic
stress changes, height growth, and fracture penetration and recession during
closure. The Placement II fracture simulator in the FracCADE software addresses
most of these effects.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 56 of 81

4.3 Calibration Decline for Fluid-Loss Behavior


An analysis and use of a specialized plot for the pressure decline during shut-in,
pioneered by Nolte, provides relationships for width, penetration, fluid efficiency and
fluid-loss coefficient in terms of rate of pressure decline and closure time after
injection. The leakoff coefficient can be quantified from the rate of pressure decline,
an important parameter in fracture treatment design. Fig. 33 shows the pressure
decline period between the end of pumping and closure.

Fig. 33. Example of fracturing-related pressures (after Nolte, 1982).

4.3.1 Review of Decline Analysis


The bases for the information within the following sections comes from SPE 25845,
“A Systematic Method for Applying Fracturing Pressure Decline” by Nolte et al and
should be consulted if required for more detail.
The assumptions of the basic decline analysis are
• constant fluid density
• constant fracture area
• constant β (dimensionless fluid pressure distribution)
• constant cf (fracture compliance)
• constant closure pressure
• constant fluid-loss area and coefficient
• Spurt loss is negligible after shut-in
• The fluid loss follows the Carter assumptions of: t dependence
• Fracture area during injection evolves with the relation A ∝ t α ( 0. 5 < α < 1).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 57 of 81

The last condition is met if injection conditions are approximately constant and
excessive height growth does not occur. The fracture area evolves in time as (see
Fig. 34)
The area exponent (α) can be found using two bounding cases:
a
A t
=  .
Af  ti 
• Lower bound η → 0, α0 = 0.5
• Upper bound η → 1,
α1 = (2n + 2)/(2n + 3) PKN
α1 = (n + 1)/(n + 2) KGD
α1 = (4n + 4)/(3n + 6) Radial
with n = the power-law fluid exponent.

Fig. 34. Schematic for fracture area and time.


The appropriate value of α for use in an application can be found by interpolating
using the actual value of efficiency;
α = 0.5 + η (α1 - 0.5).
The interpolated values of α over the full range of n are provided in Table 3, from
which a typical value is α ≈ 0.6 for 0.4 < n < 0.6 and 0.2 < η < 0.6.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 58 of 81

Table 3. Interpolated Values of α Over the Full Range of n


n′ 0.4 0.6 0.8 1.0
Efficiency PKN KGN Radial PKN KGD Radial PKN KGD Radial PKN KGD Radial
0 0.50 0.50 0.50 0.50 0.50 0.50 0.50 0.50 0.50 0.50 0.50 0.50
0.2 0.55 0.52 0.55 0.55 0.52 0.56 0.56 0.53 0.57 0.56 0.53 0.58
0.4 0.60 0.53 0.61 0.61 0.55 0.63 0.61 0.56 0.64 0.62 0.57 0.66
0.6 0.64 0.55 0.67 0.66 0.57 0.69 0.67 0.59 0.71 0.68 0.60 0.73
0.8 0.69 0.57 0.72 0.71 0.59 0.76 0.73 0.61 0.79 0.74 0.63 0.81
1.0 0.74 0.58 0.78 0.76 0.62 0.82 0.78 0.64 0.86 0.80 0.67 0.89

4.3.2 Volume Function g


For the decline assumptions, the volume lost during pumping and shut-in can be
derived analytically and expressed in terms of volume function g(∆tD).
The volume lost during pumping;
VLP = 2κgo C L rp A f t p .
The volume lost during shut-in;

[
VLS = 2C L rp A f t p g( ∆t D ) − go , ]
Where:
g( ∆t D ) = 4 / 3[(1 + ∆t D ) 3 / 2 − ∆t D3 / 2 ]
4
with go = g(o ) = for α = 1,
3
g( ∆t D ) = (1 + ∆t D )sin − 1 (1 + ∆t D ) − 1/ 2 + ∆t D
1/ 2

π 1
with go = for α = ,
2 2
Sp
κ = spurt correction = 1 + .
go C L t p
g(∆tD) as function of ∆tD is shown in Fig. 35. It is important to note that the difference
between the upper and the lower bound decreases significantly during the shut-in
period (that is, the precise value of α is not critical).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 59 of 81

Fig. 35. Dimensionless volume function for fracture closure


(after Nolte, 1986).

4.3.3 Fluid Efficiency


The fluid efficiency is the ratio of fracture volume to the total volume injected. For
ideal conditions and no proppant, the efficiency can be expressed in terms of the
dimensionless closure time (refer to Section 4.3.4 for efficiency in terms of
G function);

V f ( ∆t = 0) VLS ( ∆t = tc )
η= =
Vi VLP + VLS
g( ∆tcD ) − go
η= .
g( ∆tcD ) + (κ − 1)go

For the case of no spurt, κ = 1 and the efficiency;

g( ∆tcD) − go
η=
g( ∆tcD)

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 60 of 81

Fig. 36. Efficiency from closure time for no proppant, no spurt loss during pumping and
other ideal assumptions given in Section 4.3.1 (after Nolte, 1986).

Fig. 36 shows the relationship of η and the dimensionless closure time for the non-
propped case. The relationship is constructed in terms of upper and lower bound
(α = 1 and α = 0.5).
For determining the effect of proppant on efficiency, the proppant volume will be
expressed as the bulk proppant volume fraction;
v prop
v prop =
Vi
Where:
Vprop = proppant bulk volume
Vi = total slurry volume injected.
The expression for efficiency can be shown as follows (See Ref. 2);
Vf (V f − Vprop )(1 − v prop ) (1 − v prop )
η= = = η'
Vi  v prop   v prop 
(Vi − Vprop ) 1 −  1 − 
 η   η 

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 61 of 81

Vf is the fracture volume at the end of pumping. Denoting the apparent efficiency η’
based on closure time assuming no proppant (that is, Fig. 36) and rearrange the
previous equation, the efficiency of a propped fracture can be determined from η’;
η = v prop (1 − η' ) + η'
4.3.4 Decline Function G
The pressure decline analysis utilizes a plot of net pressure versus the
dimensionless decline function G;
4
G ( ∆t D ) = [ g( ∆t D ) − go ].
π
A combination of fracture compliance, material balance and the relationship between
the generated fracture area and time permit the development of the pressure decline
analysis. If the ideal assumptions of the basic decline analysis hold, the basic
pressure decline relation gives;
dpw π dG( ∆t D )
cf =− t p C L rp .
dt 2 dt

At shut-in, the pressure is pws with ∆tD = 0 giving a relationship between pw versus
G(∆tD);
πC L rp t p
Pw = pws − G( ∆t D ) .
2c f

This provides a straight-line interpretation with negative slope, mG = p*, of the plot of
pw versus G(∆tD) as shown in Fig. 37. Where p* is defined as;
πC L rp t p
p* = =
2c f
This relationship is used to infer a leakoff coefficient (CL). It should be noted that if
the ideal assumptions are not valid, the straight line can not be identified and mG ≠p*.
Fig. 37 shows the conceptual response of the ideal pressure decline.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 62 of 81

Fig. 37. Conceptual response of pressure decline versus Nolte time-function


(after Castillo, 1987).
The G-plot is analogous to the use of a Horner plot for pressure buildup or falloff to
characterize reservoir flow parameters. In addition to the leakoff parameters, it
provides the diagnostic for the deviations from ideal behavior, as discussed in
Section 4.3.5.
From the efficiency expression in Section 4.3.3, definition of G and go ≈π;/2, the fluid
efficiency for ideal behavior can be expressed in terms of “G” function;
Gc
η' =
2κ + Gc

where Gc = G(∆tcD) = G at closure and κ = 1 for no spurt.


The values of decline function “G” are provided in Table 4 in terms of α and ∆tD. This
table is used to calculate G(∆tD) if the DataFRAC software is not available. α is
typically between 0.5 and 0.7. For most applications, α ≈ 0.6 can be assumed with
sufficient accuracy for fracture closure analysis.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 63 of 81

Table 4. Values of Decline Function "G"


α 0.5 0.6 0.7 1.0
∆tp g η G g η G g η G g η G
0.00 1.57 .000 0.000 1.52 .000 0.000 1.48 .000 0.000 1.33 .000 0.000
0.02 1.60 .018 0.038 1.55 .020 0.039 1.51 .021 0.041 1.37 .027 0.046
0.04 1.63 .035 0.073 1.58 .038 0.076 1.54 .041 0.080 1.40 .050 0.089
0.06 1.66 .051 0.108 1.61 .055 0.112 1.57 .058 0.117 1.44 .071 0.130
0.08 1.68 .066 0.141 1.64 .070 0.147 1.60 .075 0.152 1.47 .091 0.169
0.10 1.71 .080 0.174 1.66 .085 0.180 1.62 .091 0.187 1.50 .109 0.207
0.12 1.73 .093 0.206 14.6 .099 0.213 1.65 .105 0.221 1.52 .126 0.244
9
0.14 1.76 .106 0.237 1.72 .112 0.245 1.68 .119 0.254 1.55 .141 0.280
0.16 1.78 .118 0.267 1.74 .125 0.277 1.70 .132 0.286 1.58 .156 0.315
0.18 1.80 .129 0.297 1.76 .137 0.308 1.73 .145 0.318 1.61 .170 0.349
0.20 1.83 .140 0.327 1.79 .148 0.338 1.75 .157 0.349 1.63 .184 0.382
0.25 1.88 .166 0.399 1.85 .175 0.411 1.81 .184 0.424 1.70 .214 0.463
0.30 1.94 .190 0.468 1.90 .199 0.482 1.87 .209 0.497 1.76 .241 0.540
0.35 1.99 .211 0.535 1.96 .221 0.551 1.92 .232 0.566 1.82 .266 0.614
0.40 2.04 .231 0.600 2.01 .241 0.617 1.97 .252 0.634 1.87 .288 0.685
0.45 2.09 .249 0.663 2.06 .260 0.681 2.03 .271 0.700 1.93 .308 0.754
0.50 2.14 .266 0.725 2.11 .277 0.744 2.08 .289 0.763 1.98 .326 0.821
0.55 2.19 .282 0.785 2.16 .293 0.805 2.12 .305 0.825 2.03 .343 0.886
0.60 2.23 .297 0.844 2.20 .308 0.865 2.17 .320 0.886 2.08 .359 0.949
0.70 2.32 .324 0.957 2.29 .336 0.980 2.26 .348 1.003 2.17 .387 1.071
0.80 2.41 .348 1.066 2.38 .360 1.091 2.35 .372 1.115 2.27 .412 1.187
0.90 2.49 .369 1.172 2.46 .382 1.197 2.44 .394 1.223 2.35 .433 1.299
1.00 2.57 .389 1.273 2.54 .401 1.300 2.52 .414 1.326 2.44 .453 1.406
1.10 2.65 .407 1.372 2.62 .419 1.399 2.60 .432 1.427 2.52 .471 1.510
1.20 2.72 .423 1.467 2.70 .435 1.496 2.67 .448 1.525 2.60 .487 1.610
1.30 2.80 .438 1.560 2.77 .450 1.590 2.75 .463 1.619 2.67 .501 1.708
1.40 2.87 .452 1.651 2.84 .464 1.681 2.82 .477 1.711 2.75 .515 1.802
1.50 2.94 .465 1.739 2.91 .477 1.770 2.89 .489 1.801 2.82 .527 1.894
1.60 3.00 .477 1.825 2.98 .489 1.857 2.96 .501 1.889 2.89 .539 1.984
1.80 3.14 .499 1.992 3.11 .511 2.025 3.09 .523 2.058 3.03 .560 2.157
2.00 3.26 .518 2.152 3.24 .530 2.186 3.22 .542 2.220 3.16 .578 2.322
2.20 3.38 .535 2.305 3.36 .547 2.340 3.34 .558 2.376 3.28 .594 2.481
2.40 3.50 .551 2.454 3.48 .562 2.490 3.46 .573 2.526 3.40 .608 2.633
2.60 3.61 .565 2.598 3.59 .576 2.634 3.57 .587 2.671 3.52 .621 2.781
2.80 3.72 .578 2.737 3.70 .589 2.774 3.68 .599 2.812 3.63 .633 2.924
3.00 3.83 .589 2.872 3.81 .600 2.910 3.79 .611 2.948 3.74 .643 3.062
3.50 4.08 .615 3.195 4.06 .625 3.234 4.05 .635 3.274 4.00 .666 3.392
4.00 4.32 .636 3.498 4.30 .646 3.539 4.29 .656 3.579 4.24 .686 3.701
4.50 4.54 .654 3.786 4.53 .664 3.827 4.51 .673 5.869 4.47 .702 3.994
5.00 4.76 .670 4.060 4.75 .679 4.102 4.73 .688 4.145 4.69 .716 4.272
5.50 4.97 .684 4.322 4.95 .692 4.635 4.94 .701 4.408 4.90 .728 4.538
6.00 5.16 .696 4.573 5.15 .704 4.617 5.14 .713 4.661 5.10 .738 4.793
6.50 5.35 .707 4.816 5.34 .715 4.860 5.33 .723 4.905 5.29 .748 5.038
7.00 5.54 .716 5.050 5.52 .724 5.095 5.51 .732 5.140 5.48 .757 5.275
7.50 5.71 .725 5.276 5.70 .733 5.322 5.69 .741 5.367 5.66 .764 5.504

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 64 of 81

4.3.5 Non-Ideal Behavior


The assumptions of the basic decline analysis are seldom met in practice. Several
deviations from these ideal assumptions that have to be considered are
• change in fracture penetration after shut-in
• height growth
• pressure-dependent fluid loss
• spurt
• closure pressure change
• compressible fluids.

4.3.5.1 Change in Fracture Penetration After Shut-in


A varying fracture length after shut-in will affect the basic pressure decline analysis.
The effect of length change is illustrated in Fig. 38, which shows an initial slope
greater than the slope at closure. The early rapid decline results from fluid flow past
the fracture tip at shut-in, that is, fracture extension after shut-in. The fracture length
will then recede toward the wellbore during closing. The slope will decrease as high
leakoff area is lost. Fig. 38 shows the G-plot for this case, which clearly indicates a
significant reduction in slope with time.
The correct value of leakoff can be obtained by applying a correction to the standard
“G” plot.
Corrected slope m′G = fcmG,
Where:
fc = correction factor
mG = slope of the G-plot near closure.
β'
fc = 1 + ∆t D f ( ∆t D )
βs
f ( ∆t D ) = 2[(1 + ∆t D )1/ 2 − ∆t D
1/ 2
] for α = 1
f ( ∆t D ) = sin − 1 (1 + ∆t D ) − 1/ 2 for α = 1 / 2

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 65 of 81

Values of fc are provided in Table 5.


The features of the G-plot also indicates that the correct value of p* can be obtained
dA f
at the transition between extension and recession, that is, when: = 0 at the end
dt
∆pw
of extension. The transition from extension to recession occurs at ≈ 3 / 4 (with
∆ps
∆ps = net pressure immediately after shut-in) and resulted in suggestion of a “3/4 rule”
to eliminate the effect of penetration changes during shut-in; that is, select p* as the
value of mG at ∆pw = 3/4 ∆ps as shown in Fig. 38.

Fig. 38. Penetration change during shut-in (after Nolte, 1990).

4.3.5.2 Height Growth


Height growth will reduce the rate of pressure decline during initial shut-in. During
this period, the decreasing height dispels fluid into the primary fracture and creates
an equivalent flow rate source. This flow rate source delays the closure time. This
period will last until the net pressure decreases to about 0.4∆σ for the barrier and
provides a height growth diagnostic (Fig. 39). The transition to a greater decline
after the height growth closes provides the period of the decline analysis similar to
the case of no growth. Applying the correction to the slope near closure provides
correct estimate of p* (the similar correction for the case of length recession). Fig.
39 also suggests the governing stress barrier difference can be defined from the G-
plot;

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 66 of 81

∆pknee
that is, ∆σ = with ∆pknee the net pressure at the “knee” of the dog-leg.
0.4

Fig. 39. Diagnostic for height growth from decline data (after Nolte, 1990).

4.3.5.3 Pressure-Dependent Leakoff


There are two mechanism for the pressure-dependent fluid loss.
1. stress-sensitive fissure
2. filtrate - and reservoir-control mechanism
Stress-Sensitive Fissures
For the stress-sensitive fissures, the governing pressure is the difference between
principal horizontal stresses (∆σH, see Section 4.2.2.2). During a fracture treatment,
the pressure within the fissure increases as fluid leaks off into it. The effective
normal stress on the fissure decreases and its permeability increases. The
permeability and fluid loss are enhanced throughout the treatment, with the effect
accelerating as the pressure increases. If the pressure continues to increase, the
pressure in the fissure can become greater than the normal stress, the fissure will
open and the leakoff is further accelerated. At this time, the pressure capacity is
reached and regulates the pressure at a constant value.
The diagnostic for the stress-sensitive fissure as shown in Fig. 40 is a progressively
decreasing log-log slope on the Nolte-Smith plot during pumping until the pressure
capacity is reached (that is, showing constant pressure condition) and a
progressively decreasing slope on the G-plot (positive curvature) during decline

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 67 of 81

period. Height growth can also produce a constant pressure condition similar to
fissures; however, height growth (Fig. 39) provides negative curvature (increasing
slope) on the G-plot; the distinguishing diagnostic for fissures relative to height
growth.

Fig. 40. Diagnostic for stress sensitive fissures from injection and decline
(after Nolte, 1990).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 68 of 81

The pressure behavior of stress-sensitive fissures is more complex than filtrate- or


reservoir-controlled fluid loss. Numerical simulations are required to properly asses
this effect. The value of the leakoff coefficient is best determined by pumping
pressure history match, using an appropriate fracture simulator. To approximate CL
∆pw
from the G-plot, the slope of the G-plot is selected at ≈ 3 / 4 that is, when the
∆ps
area is constant and with a high value of net pressure so the effect of fissures
dominates the fluid loss.
Filtrate- and Reservoir-Controlled Mechanism
If leakoff is controlled by the filtrate viscosity (Cv) or by reservoir permeability and
compressibility (Cc), the leakoff is pressure dependent. No significant pressure
dependence is expected for a wall-building fluid. For the filtrate- and reservoir-
control fluid loss, the governing pressure is the difference between fracturing
pressure (pf) and reservoir pressure (pr). This mechanism is significantly affected by
pressure change during decline as shown in Fig. 41. The figure shows the G-plot
found by numerical simulation and indicates a significant reduction in slope with time,
that is, mG < p* at later stage of closure. Using the fracture simulator, the leakoff
coefficient was found to be approximated by the slope of the net pressure at
∆pw
≈ 3 / 4 as shown in Fig. 41.
∆ps

Fig. 41. Decline analysis for filtrate and reservoir control leakoff
(after Nolte, 1993).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 69 of 81

4.3.5.4 Spurt
Spurt (Sp) is the fluid volume lost during the formation of a filter cake. It is negligible
for low-permeability formations (less than one md). The rate of spurt is controlled by
the filtrate (Cv) and reservoir fluid (Cc) during the spurt period (tsp). Generally, spurt is
negligible after shut-in since the new area exposed is small; therefore, spurt loss has
a negligible effect on the pressure decline and cannot be defined from the pressure-
decline analysis.
For large values of spurt (which can occur at high ∆p or in high-permeability
formations without using effective fluid-loss control additives), the large spurt time
and volume can affect the decline analysis. Spurt will affect the pressure-decline
analysis based on the “3/4” rule if the spurt time is greater than the time at
∆pw
≈ 3 / 4 . It can be shown from the volume balance relation at shut-in that the
∆ps
spurt will not affect the decline analysis if;
tsp κη
< .
ti 4(1 − η)
Although the spurt generally does not influence the decline analysis using the “3/4”
rule, the magnitude of spurt during injection may be important and must be
characterized. The effect of spurt can be eliminated for the treatment by performing
a calibration treatment of a size comparable to the treatment and using effective
fluid-loss control additives. The spurt can be estimated from the laboratory-
Sp
determined ratio of for representative formation and fluid samples. Assuming
CL
the ratio is the same in the laboratory and field, Sp can be defined by using CL from
Sp
calibration treatment and keeping the ratio of constant;
CL
Sp
Spurt correction κ = 1 +
go C L ti

The spurt correction κ is used to account for additional fluid loss due to spurt. The
effective fluid-loss coefficient during pumping = κCL.

4.3.5.5 Closure Pressure Change


The closure pressure is assumed constant in the basic pressure-decline analysis.
Leakoff of the fracturing fluid under high pressure will cause an increase of the
closure pressure because of poroelastic effect. Although not common, evidence of
increased closure pressure with injection time has been reported in the field. The
poroelastic stress changes have little effect on fracture geometry, but the pressure is
increased by an amount approximately equal to the stress change during injection or
decline.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 70 of 81

The most significant stress changes occurs for the PKN model. The upper bound of
the stress change for this case;

2.64 MC L t
∆σ ≈
φct h
1/ 4
 
 
M ≈ 1 − 0.6  kt 
 2
 φµct  h  
  2  
The equation indicates that the stress change is proportional to the fluid-loss
coefficient, and hence depends on the controlling fluid-loss mechanism, that is, wall
cake (Cw), filtrate viscosity (Cv) or reservoir control (Cc). The analysis by Nolte has
shown that this closure stress change is bounded by a maximum for constant CL (Cw
controlled or small ∆pf) and a minimum for Cc. Fig. 42 shows the wellbore stress
change for Cc = 0.001 ft / min and 0.002 ft / min.

Fig. 42. Stress change during injection/shut-in for Cc (after Nolte et. al., 1993).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 71 of 81

The rate of change of ∆σ;


d∆σ 1.32 M 2C L
≈ .
dt φct h t

The rate of stress change decreases with time. When stress changes are
significant, the closure pressure test should be conducted immediately after the
decline of the calibration treatment instead of before the treatment.
Fig. 42 indicates that the stress change is generally positive; the stress continues to
increase during pumping and after shut-in. This would result in an underestimate of
CL. A conservative estimate of CL can be obtained from the inferred CL as follows;

 0.43 
C' L = C L 1 +
 φct E' 

4.3.5.6 Compressible Fluids


Most fracturing fluids can be assumed to have constant density since they are
relatively incompressible compared to the elastic response of the fracture. The basic
pressure-decline analysis assumes that the fluid density is constant
(incompressible). Foams, however, are much more compressible than liquids. The
foam density change can have an effect on the pressure-decline analysis and must
therefore be considered.
When the fracturing fluid is compressible and significant warming of the fracturing
fluid occurs after shut-in, the thermal expansion of a compressible fracturing fluid will
affect the pressure behavior during decline. The analysis indicates that the pressure
of a compressible fluid will decline at a slower rate compared with an incompressible
fluid (decrease the slope of the G-plot). The fluid-loss coefficient will therefore be
underestimated.
The effect of foam fluids on the decline analysis is expressed in terms of the ratio of
∆Vg ∆Vg
the change in gas volume to fracture volume, denoted as . The value of
∆V f ∆V f

is governed by the effect of temperature and pressure of the compressible fluid.


Fig. 43 shows a simulation for an extreme case (foam fracturing of hot, shallow
formations) using a fracture simulator; a case of high temperature, high efficiency
and low closure pressure to illustrate the large effect on the gas volume increase.
For this case the error is about 13% for the fluid-loss coefficient obtained by the “3/4”
rule.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 72 of 81

Although the effect for the increase in gas volume of the compressible fluid during
shut-in is small, a correction factor can be applied to the inferred CL. The correction
term is provided in SPE 25845 (Nolte et al, 1993) and is a relatively complex relation.

Fig. 43. Relative volume change of gas


(after Nolte et. al., 1993).

4.3.6 Fluid Efficiency Based on Pressure Analysis


The expressions for fluid efficiency in Section 4.3.3 and Section 4.3.4 define this
parameter in terms of dimensionless closure time for ideal behavior with or without
spurt. These expressions can be generalized for non-ideal behavior by defining the
efficiency in terms of shut-in net pressure (∆ps) and match pressure (p*). The fluid
efficiency can be expressed as follows;
Vf Vf Af w
η= = = .
Vi VLP + V f 2κgo C L rp A f t p + A f w
From the relationship of width and pressure directly after shut-in (w = cf ∆ps), and
2c f p *
CL = ,
πrp t p

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 73 of 81

A f c f ∆ps
η= ,
4
κA c p * go + A f c f ∆ps
π f f

π
with go ≈ ,
2
∆ ps
η≈ ,
2κp * + ∆ps
G*
η≈ ,
2κ + G *
∆ps
where G* = . For ideal conditions, G* = Gc and the expression for efficiency
p*
becomes as in Section 4.3.4. For non-ideal behavior, p* ≠ mGC and G* ≠ Gc. Because
the ideal assumptions are not generally valid, the closure point is not a reliable
predictor for efficiency.

4.3.7 Decline-Analysis Procedure


As discussed in the previous section, the corrections can be provided for violating
the basic assumptions. The non-ideal behavior is related to penetration changes,
height growth, pressure-dependent fluid loss, spurt, closure pressure change, and
density effects. The decline-analysis procedure is (see Fig.44):
∆pw
1. Find the slope of the “G” plot (mG) at ≈ 3 / 4 (that is, m3/4, referred to as the
∆ps
“3/4 rule”).
2. For the KGD and radial models,
p* = m3/4
3. For the PKN model,
p* = max of { m3/4 , m′G }
m’G = fc mGc
where fc is the correction factor (see Table 5) and mGc is the slope near closure.
4. Calculate CL.

2c f p *
CL =
πrp t p
5. Calculate η.
G* ∆p
η= where G* = s
2κ + G * p*

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 74 of 81

β′
The correction factor (fc) is the product of 1 + ∆t D f ( ∆t D ) and is provided in
βs
Table 5. From numerical simulations, β′≈1.
The decline-analysis procedure provides the following results.
• For height growth without pressure-dependent fluid loss,
p* ≈ m′G,
In this case, m′G > m3/4.
• For no significant height growth but pressure-dependent loss,
p* ≈ m′3/4,
In this case, m′G > m3/4.

Fig. 44. Decline analysis using “¾” rule (after Nolte, 1990).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 75 of 81

Table 5. Correction Factors fc As Function Of ∆tD


∆ tD fc ∆tD)
G(∆ ∆ tD fc ∆tD)
G(∆
0.00 2.09 0.00 2.10 1.42 2.23
0.10 1.77 0.17 2.20 1.41 2.31
0.20 1.68 0.33 2.30 1.41 2.38
0.30 1.63 0.47 2.40 1.41 2.45
0.40 1.59 0.60 2.50 1.41 2.53
0.50 1.56 0.72 2.60 1.40 2.60
0.60 1.54 0.84 2.70 1.40 2.67
0.70 1.52 0.96 2.80 1.40 2.74
0.80 1.50 1.07 2.90 1.40 2.80
0.90 1.49 1.17 3.00 1.40 2.87
1.00 1.48 1.27 3.10 1.39 2.94
1.10 1.47 1.37 3.20 1.39 3.00
1.20 1.46 1.47 3.30 1.39 3.07
1.30 1.46 1.56 3.40 1.39 3.13
1.40 1.45 1.65 3.50 1.39 3.19
1.50 1.44 1.74 3.60 1.39 3.26
1.60 1.44 1.83 3.70 1.39 3.32
1.70 1.43 1.91 3.80 1.38 3.38
1.80 1.43 1.99 3.90 1.38 3.44
1.90 1.43 2.07 4.00 1.38 3.50
2.00 1.42 2.15

4.3.8 Steps to Correct Decline Analysis Using the FracCADE Software


The following steps are required to correctly analyze the pressure decline from a
calibration test:
1. Ensure the decline data is good.
After the data has been imported into the DataFRAC software, examine the
G-plot using 'Graphics' mode to see if the data are smooth and continuous.
Remove any bad initial data which included the pressures when the pumps were
still rolling over. This initial data gives high net pressure and steep initial decline.
Another type of bad data is not enough data points (resembling a dot-to-dot
puzzle).

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 76 of 81

2. Determine the analysis type to be performed on the “G” plot.


For the basic decline analysis when the ideal assumptions hold, the linear
portion of the “G” plot can be determined using 'Manual' analysis. For the
condition of non-ideal behavior, the procedure of the “3/4” rule as discussed in
the previous section can be performed using “recession” analysis in the
DataFRAC software.
3. Match net pressures.
The last step is to match the net pressure at shut-in with the predicted net
pressure from the fracture simulator using the parameters derived from the
analysis of the “G” plot. The match is performed by adjusting Young's modulus
(E) or gross fracture height (H) for the PKN model and adjusting E or rock
toughness (K) for the KGD or radial model. This pressure match assumes ideal
behavior and application of the ideal models. For non-ideal behavior, the net
pressure match should be done using the Placement module in the DataFRAC
software.

4.3.8.1 The DataFRAC Software


The well, reservoir, rock mechanics, and fluids screens should be completed before
performing the DataFRAC analysis. The Job Record Data Entry form is used to read
the pressure datafile into the FracCADE software (refer to the FracCADE User's
Manual for complete information).
The fluid Sp and Cw values can be obtained from laboratory data or the Fracturing
Materials Manual  Fluids. The ratio of Sp and Cw will be used and kept constant in
calculating the new spurt based on the leakoff coefficient (CL) obtained from the
analysis. The new values of Sp and Cw are used to determine efficiency.
The DataFRAC analysis information should be completed to define how the analysis
of the G-plot is controlled and performed. The types of analysis available in the
DataFRAC program are
1. Graphic Analysis
The 'Graphic' option allows the user to look at the G-plot. It is normally used first
to determine if the data set is good. This option does not perform any
calculation.
2. Automatic Analysis
The 'Automatic' option requires the least amount of interaction from the user. It
is used for the ideal conditions of the basic pressure decline by automatically
selecting the best straight line of the G-plot. The program uses the derivative of
the G-plot to look for the minimum curvature within the analysis range. The
straight line selected likely will not correspond to p* and may or may not
correspond to the fracture closure period.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 77 of 81

3. Manual Analysis
The 'Manual' option is also used for the ideal conditions. The user supplies the
maximum and minimum pressure constraints to be used for determining the
straight line of the G-plot.
4. Recession Analysis
The recession analysis with “closure pressure” as the closure variable is the
preferred option and uses the “¾” rule or correction to slope on the G-plot as
outlined in the previous sections to account for the non-ideal behavior.
Closure Variable
Closure input variables control the determination of the closure point, which indicates
if closure pressure or closure time is to be specified. Using “closure pressure”, the
corresponding “closure time” will be extrapolated and vice versa. The closure
pressure should be determined from an independent closure test analysis (see
Section 4.1). If the closure variable is not specified, the program will automatically
take the last point on the straight line of the G-plot as the closure point.

4.3.8.2 G-plot Interpretation by the DataFRAC Software


∆pw
The value of Yint (Y intercept from the tangent line to the G-plot at ≈ 3 / 4 or at
∆ps
closure) compared to the actual Initial shut-in pressure (ISIP) gives some indications
of the fracture evolution. The difference should generally not be greater than 200
psi.
A condition in which Yint is less than the ISIP is equivalent to a positive curvature of
the G-plot due to the non-ideal behavior described in the previous sections, that is,
penetration change and pressure-dependent fluid loss.
When Yint is greater than the ISIP, it indicates a negative curvature of the G-plot
because of height growth into stress barriers. It can also be concluded that the
equivalent 2D model is PKN since height growth into stress barriers is inconsistent
with the basic requirements of the radial or KGD model.
The inequality of Yint and the ISIP is corrected by shifting the tangent line upward or
downward to intercept the ISIP. This will also allow the user to use a proper value of
net pressure based on the ISIP instead of Yint for determining the efficiency and
making a net pressure match with the fracture simulator. Answer “Yes” to the
“Adjustment for Initial Deviation” field to perform this function.

4.3.8.3 Modulus, Height, or Fracture Toughness Calibrations


The most important and uncertain parameters for a proper evaluation of the leakoff
coefficient are Young's modulus, total fracture height, and fracture toughness. The
values of Young's modulus obtained from the log can be crossed-checked and

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 78 of 81

calibrated using the DataFRAC software. The fracture height for the PKN model and
the fracture toughness for the KGD or radial model can also be calibrated.
The procedure is to make the simulator net pressure match up with the analysis by
adjusting the Young's modulus and holding other parameters constant. The gross
fracture height for the PKN model and the fracture toughness for the KGD or radial
model can also be adjusted until the net pressure from the fracture simulator
matches that from the analysis.
For the case with barriers, the preferred method is to select the height as the gross
pay (or sand) section and calibrate the modulus from the DataFRAC analysis. The
calibrated modulus can be much greater (that is, twice) than actual modulus to
account for interbedded shales (that is, high stress zone).
Compliance for each model gives an understanding as to which parameter is
controlling the pressure response. The compliance for each model is

βh f
• PKN: cf ∝ E′ = plain strain modulus
E′

• KGD: 2βx f E′ = E/1 - v2


cf ∝
E'
βR
• Radial: cf ∝
E'
The average width is expressed in terms of fracture compliance (cf) and net wellbore
pressure (∆pf), that is, w = cf ∆pf. This indicates that both total fracture height and
Young's Modulus have a large impact on net pressure for the PKN model. For the
KGD and radial models, Young's Modulus has an impact on pressure. The fracture
width and length (xf or R) are also affected by fracture toughness for the KGD and
radial model. The fracture toughness will therefore have an important effect on net
pressure for these models.
In a 1991 publication, Nolte incorporated toughness in the PKN model. The β factor,
and therefore, the pressure is affected by the toughness. The pressure matching
using toughness for the PKN model has not been implemented in the DataFRAC
software because the PKN pressure is not very dependent on the tip behavior.

4.3.8.4 The β Ratio


The β ratio is the ratio of the average net pressure in the fracture and the wellbore
net pressure;
∆p f p f − pc
β= =
∆p f pw − pc

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 79 of 81

This parameter is necessary for volumetric calculations of the fracture in terms of the
pressure value at the wellbore. The analysis of fluid flow in the fracture indicates
there is a gradient of pressure from the maximum value at the well, pw, to the
minimum value at the fracture tip, approximately pc. Fig.45 shows an example of the
pressure and flow profiles during pumping and after shut-in using the PKN model.

Fig. 45. Pressure and flow rate in fracture before and after shut-in
(after Nolte, 1986).
The value of β during pumping (βp) differs from the value of β after shut-in (βs).
During pumping,

n' +2 PKN
βp ≈
n' +3
β p ≈ 0.9 KGD and Radial
During shut-in,
2n' +2 PKN
βs ≈
2n' +3
β s ≈ 0.95 KGD and Radial
βs and βp are used to convert the net pressure at shut-in (∆ps) to the net pressure at
the end of pumping (∆pp);
β
∆p p = s ∆ps .
βp

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
DataFRAC Service Dowell
Page 80 of 81

The net pressure at the end of pumping is used by the DataFRAC program in the net
pressure matching to calibrate Young's modulus, total fracture height, or fracture
toughness.

4.3.9 Post Proppant Fracture Analysis


The “G” plot analysis after the propped fracture treatment can provide information on
the effectiveness of the fracture treatment. The effectiveness is indicated by the
ratio of net pressure at closure on proppant to net pressure at shut-in;
w prop ∆p prop
=
whyd ∆ps
If fracture closure did not occur until most of the net pressure was lost (that is,
∆pf = 0), very little of the fracture width was propped and hence, the job was not
effective.

Fig. 46. Diagnostic for closing on proppant from decline data


(after Nolte, 1990).
The closure on proppant will change the rate of pressure decline, that is, with
significant slope change. Fig.46 illustrates the effect of fracture closing on proppant.
Two cases affect the pressure decline behavior after closure.
1. Decrease in rate of pressure decline. A decrease in the rate of pressure decline
is indicative of a relatively impermeable proppant-pack caused by unbroken fluid
and the filter cake impairing communication between the fracture and the
wellbore.

DOWELL CONFIDENTIAL
Section 700.1
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
DataFRAC Service
Page 81 of 81

2. Increase in rate of pressure decline. An increase in the rate of pressure decline


indicates the fracture system stiffening during closure on a permeable pack
communicating freely with the well after closure.
Although the initial closure period may be relatively free of the effects of proppant,
the G-plot may be different from the representative condition and any analysis could
be subject to a high degree of uncertainty. Therefore, an unpropped calibration
treatment is always preferred over a propped fracture pressure decline; but for
obtaining some insight into a failed treatment, the decline from the treatment can be
used to infer the fluid-loss coefficient. The effect of proppant or the inference of
efficiency is discussed in Section 4.3.3.

4.3.10 References
Comprehensive discussion of fracturing pressure analysis are provided in the
following publications:
1. Nolte, K.G.: Fracturing Pressure Analysis, Recent Advances in Hydraulic
Fracturing, J. Gidley et al. (eds.), Monograph Series, SPE, Richardson, TX
(1989) 12, Chap. 14.
2. Nolte, K.G. and Economides, M.J.: Fracturing Diagnosis Using Pressure
Analysis, Reservoir Stimulation, second edition, Prentice Hall, Englewood Cliffs,
NJ (1989) Chap. 7.
3. Nolte, K.G.: “Fracturing Pressure Analysis for Nonideal Behavior,” JPT
(Feb., 1991) 210-18.
4. Nolte, K.G.: “A General Analysis of Fracturing Pressure Decline With Application
To Three Models,” SPEFE (Dec. 1986) 571-83.
5. Nolte, K.G., Mack, M.G. and Lie, W.L.: “A Systematic Method For Applying
Fracturing Pressure Decline, Part 1,” SPE 25845, Denver (April 1993).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 1 of 30

FOAM FRACTURING

1 Introductory Summary............................................................................................................. 2
1.1 Foam Properties .................................................................................................................. 3
1.2 Foam Types ......................................................................................................................... 3
1.3 Foam Stability ...................................................................................................................... 4
1.4 Applications.......................................................................................................................... 4

2 Design ....................................................................................................................................... 4
2.1 Choosing a Foam................................................................................................................. 4
2.1.1 The Liquid Phase........................................................................................................ 5
2.1.1.1 Linear Polymers ............................................................................................... 5
2.1.1.2 Crosslinked Polymers ...................................................................................... 6
2.1.1.3 Hydrocarbons and Alcohols ............................................................................. 7
2.1.2 The Gas Phase........................................................................................................... 7
2.1.2.1 Gas Behavior ................................................................................................... 9
2.1.2.2 Gas Solubility ................................................................................................. 10
2.1.3 Foaming Agent Selection.......................................................................................... 11
2.1.3.1 Material Compatibility with Foaming Agents .................................................. 11
2.2 Foam Rheology.................................................................................................................. 12
2.3 Fluid-Loss Properties ......................................................................................................... 13
2.3.1 Two-Phase Behavior of the Foam ............................................................................ 13
2.3.2 Wall-Building Effects................................................................................................. 13
2.4 Conductivity Damage ......................................................................................................... 14
2.5 Foam Quality...................................................................................................................... 16
2.6 Foam Texture..................................................................................................................... 19
2.7 Proppant Compensation .................................................................................................... 19
2.7.1 No Proppant Compensation ..................................................................................... 20
2.7.2 Constant Bottomhole Quality .................................................................................... 21
2.7.3 Decreasing Bottomhole Quality ................................................................................ 21
2.7.4 Constant Internal Phase ........................................................................................... 22

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 2 of 30

2.8 Friction Pressure ................................................................................................................22


2.9 Yield Stress ........................................................................................................................23
2.10 Limitations of Application..................................................................................................23
2.11 Job Design........................................................................................................................23
2.12 Calculations ......................................................................................................................23
2.12.1 Pressures and Rates...............................................................................................23
2.12.2 Equipment Requirements........................................................................................25
2.12.3 Material Requirements ............................................................................................26

3 Execution ................................................................................................................................26
3.1 Foam Generation ...............................................................................................................28
3.2 Material Balance.................................................................................................................30
FIGURES
Fig. 1. The effect of polymer loading on foam viscosity (50% quality foam).................................5
Fig. 2. The effect of various polymers on foam stability. ..............................................................6
Fig. 3. The effect of foam quality on viscosity (StableFOAM fluid). ............................................12
Fig. 4. Leakoff of a foam into the rock matrix. ............................................................................13
Fig. 5. Dimensionless polymer concentration factor...................................................................15
Fig. 6. Polymer concentration versus proppant-pack retained permeability...............................16
Fig. 7. Bubble arrangements for various foam-quality ranges....................................................17
Fig. 8. Proppant concentration limits in foam fluids. ...................................................................18
Fig. 9. The effect of proppant compensation methods on bottomhole foam quality. ..................20
Fig. 10. Friction through perforations. ........................................................................................25
Fig. 11. Schematic of foam fracturing treatment.........................................................................27
Fig. 12. Laminar and turbulent flow areas of foamed fluids........................................................29
Fig. 13. Foam generator.............................................................................................................29
TABLES
Table 1. Summary of Foam Fracturing Fluids ..............................................................................3
Table 2. Comparison of Nitrogen and Carbon Dioxide.................................................................7
Table 3. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant ........................15

1 Introductory Summary
A foam fracturing fluid is a stable emulsion composed of a liquid (external or
continuous) phase surrounding a gas (internal, dispersed, or non-continuous) phase
and a surfactant (foaming agent).
Foam fracturing fluids are characterized by their “quality.” The quality of a foam (Q)
is defined as the ratio of gas volume to the liquid and gas volume.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 3 of 30

Vg
Q=
Vg + VL

Vg is the volume of the gas-phase and VL is the volume of the liquid-phase. The
liquid ratio of the total foam is then 1 - Q.
Section 2.5 provides a discussion on foam quality.

1.1 Foam Properties


Foam fracturing fluids, as compared to nonfoam fluids, are particularly well suited for
fracturing because of some very unique properties. These include:
• stored compressed gas for better cleanup
• good fluid efficiency
• low fracture conductivity damage
• equivalent rheological performance at reduced polymer loading.

1.2 Foam Types


The most common and most versatile types of foams are aqueous-base foams
containing a polymer in the liquid phase and nitrogen as the gas phase. Foams
containing alcohol, oil or carbon dioxide are used to improve performance based on
a specific requirement. A summary of foam fracturing fluids is provided in Table 1.

Table 1. Summary of Foam Fracturing Fluids


Name Liquid Phase Advantages
StableFOAM Water or Polymer Solution Good overall use, easy to mix, cost
effective, good rheology, good fluid-
loss properties and stability.
SuperFOAM Crosslinked Polymer High viscosity, stable at low-foam
qualities for higher proppant
concentrations or higher hydrostatic
pressures.
Alcoholic 20% to 40% Alcohol Less retained water on formation,
and better cleanup particularly in dry,
low-water formations.
Stable-Oil- Foam Gelled or Ungelled Hydro- No water for water-sensitive
carbon formations.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 4 of 30

1.3 Foam Stability


Foam stability is critical to the foam performance in the fracture. Stability maintains
the dispersion of the gas in the liquid which in turn controls the rheology and fluid-
loss properties of the foam. Factors affecting stability are:
• surfactant type
• surfactant concentration
• foam quality
• polymer type and concentration
• mixing energy.
The stability of a foam is normally measured under static conditions at low
temperature and pressure. These laboratory tests do not represent the actual
downhole stability and only serve as a guide to determine which polymer or
surfactant offers the most stability.

1.4 Applications
Foam fracturing fluids perform best in the following applications.
• depleted or underpressured wells
• water-sensitive formations
• low-permeability gas wells.

2 Design
A foam is a stable dispersion of a gas in a liquid. An unstable dispersion is also a
foam, but only for a short period of time. Once segregation of the phases occurs,
the properties of the foam also disappear and the fluid becomes only an energized
fluid. Three conditions are necessary to create a stable foam.
• A foaming surfactant at sufficient concentration and free of contaminants must be
used.
• The liquid and gas must be in the proper ratio. Segregation between the liquid
and gas phases will readily occur if an insufficient quantity of gas is present. The
foam may invert to a mist with the gas as the outside phase if too much gas is
present.
• The mixing energy must be sufficient to create the foam.

2.1 Choosing a Foam


Starting with the simplest, most versatile foam is the best approach when choosing a
foam fluid for a specific application. If a deficiency exists because of a limitation in

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 5 of 30

the first choice, then a modified foam fluid that addresses the deficiency is more
appropriate.

2.1.1 The Liquid Phase


Water without a polymer is not commonly used as the liquid phase because of
limited stability. Enhanced stability can be achieved by adding a polymer, increasing
the polymer concentration or by crosslinking the polymer. While crosslinked
polymers exhibit the greatest stability because of their high viscosity, linear polymers
are also very good at stabilizing the foam.

2.1.1.1 Linear Polymers


The presence of a polymer in the liquid phase of a foam substantially increases the
foam viscosity. The foam stability is directly dependent on the viscosity of the liquid
phase. The greater the viscosity, the less drainage of liquid from the bubble and
therefore the more stable the foam. Enhanced stability of the foam results in much
more efficient and reliable proppant transport.
Fig. 1 illustrates the effect of polymer loading on foam viscosity. A foam without a
polymer exhibits very low viscosity (less than 5 cp at 175°F [79°C]). The need for
polymer in the foam is even more evident as temperature increases.

Fig. 1. The effect of polymer loading on foam viscosity (50% quality foam).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 6 of 30

Polymer Types
Fig. 2 illustrates the effect of various polymers on foam stability. Guar,
hydroxypropylguar (HPG), hydroxyethylcellulose (HEC) and xanthan gum are the
most commonly used polymers.
The half-life of a foam is the time necessary for one-half of the liquid phase to break
out of the foam under atmospheric conditions. Half-life measurements are used only
as qualitative indicators of foam stability in the laboratory. Foam half-life is much
longer in the fracture under high-pressure conditions. A 75% quality foam without
polymers generally yields a half-life of less than five minutes. Addition of polymers
increases the half-life of the foam substantially.
Of the available polymers, xanthan gum is the most efficient and can be used at
much lower concentrations than other polymers.

Fig. 2. The effect of various polymers on foam stability.

2.1.1.2 Crosslinked Polymers


A crosslinked polymer in the liquid phase allows the gas content to be decreased
which results in a larger hydrostatic pressure and lower surface treating pressure.
However, the viscosity of the crosslinked liquid phase is greater than the linear liquid

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 7 of 30

phase which increases the friction pressure and may reduce some of the benefits of
the lower treating pressure.
The major advantage of using a crosslinked foam is the ability to achieve higher
proppant concentrations in the fracture. Assuming the maximum proppant
concentration at the blender is 20 PPA, the maximum proppant concentration in an
uncrosslinked 70% quality foam is approximately seven pounds proppant added
(PPA) because all of the proppant must be added to the liquid phase (30% of the
foam). Because the gas content is lower in a crosslinked foam, typically a 30%
quality foam, the maximum proppant concentration can be higher (15 PPA).
Treatment design may allow a lower foam quality in later stages of a treatment to
achieve higher proppant concentrations.

2.1.1.3 Hydrocarbons and Alcohols


Hydrocarbon foams containing carbon dioxide are impractical due to the high
solubility of the carbon dioxide in the oil. Hydrocarbons foamed with nitrogen are
costly fluids because of the fluorocarbon surfactants required to make a stable foam.
Alcohol (chiefly methanol) foams can be used in dry gas reservoirs to prevent
relative permeability problems. Specific limits for the alcohol content in the aqueous
phase are provided in the Dowell Location Safety Standards.

2.1.2 The Gas Phase


Nitrogen and carbon dioxide are the gases most commonly used in foam fluids.
Formation characteristics, fluid compatibility and economics are major factors that
are considered during the decision-making process. Nitrogen is an inert gas, and is
the most frequently used because it is versatile. Carbon dioxide is more soluble in
water than nitrogen so more carbon dioxide is required to saturate the liquid and
create the foam. Table 2 provides a comparison of nitrogen and carbon dioxide.
Table 2. Comparison of Nitrogen and Carbon Dioxide
Property Nitrogen Carbon Dioxide
Hydrostatic Head Low High
Reactive Inert Yes
Solubility in Water Low Moderate
Solubility in Oil Low High
Surface Tension Reduction None Good
Compressibility High Low
Temperature 100°F (38°C) 20° to 40°F (-7° to 4°C)

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 8 of 30

In certain applications, carbon dioxide may have an advantage. These include:


• greater hydrostatic pressure (liquid carbon dioxide is more dense than nitrogen)
resulting in lower treating pressure
• more expansion during flowback (aids in total fluid recovery)
• may prevent or remove water blocks.
The solubilized portion of carbon dioxide reduces the interfacial tension of the
fracturing fluid to levels as low as those obtainable by many surfactants. Carbon
dioxide has an advantage in that the carbon dioxide is soluble in the water whereas
the surfactant may loose its efficiency by absorbing onto the rock surfaces. This
becomes more critical in tight and low-pressure formationsprimary candidates for
foam fracturing fluids.
Carbon dioxide is extremely soluble in oil. Stable-Oil-Foams normally use nitrogen
as the gas phase when the foam quality is greater then 50%. Fluids that are
saturated with carbon dioxide have low interfacial tension which reduces capillary
pressure and damage. Carbon dioxide also reduces the viscosity of formation oils
and gives an initial production “kick.” This is of little consequence to long-term
production but gives the appearance of high production. However, at the same time
care must be taken to ensure that the carbon dioxide and the relatively large
proportion of surfactants pumped into the formation do not create emulsions that
could damage the permeability and reduce productivity.
Carbon dioxide, unlike nitrogen, is not compatible with all liquid phases. Carbon
dioxide is not recommended in the following fluids.
• YF100, YF100D, YF200, YF200D, YF500HT, YF600LT, and YF600HT. Carbon
dioxide will interfere with the crosslinking mechanism by lowering the fluid pH
value.
• Stable-Oil-Foam. Carbon dioxide is highly soluble in Stable-Oil-Foam and will
reduce the viscosity of the hydrocarbon.
Carbon dioxide is easily dispersed in the YF300LPH and YF400LPH series of fluids.
The pH buffer contained in the crosslinker solution maintains a constant fluid pH
value of approximately 4, which simulates a saturated carbon dioxide environment.
Consistent fluid performance is ensured despite variations in the carbon dioxide
concentration or complete loss of carbon dioxide during job execution.
Carbon dioxide is pumped at the wellhead in liquid form. The mixture of aqueous
fluid and liquid carbon dioxide, although not a foam by definition, forms a two-phase
emulsion which has properties similar to foam.
The critical temperature (triple point) of carbon dioxide is approximately 88°F (31°C).
Carbon dioxide is a supercritical fluid commonly referred to as a gas at temperatures
greater than 88°F (31°C). The transition from liquid to supercritical fluid does not
affect the physical properties of either the carbon dioxide or the foam provided the
treating pressure is greater than 1080 psi, the critical pressure of the carbon dioxide.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 9 of 30

Halliburton Services holds patents for exclusive use of 53% to 96% quality foams
made with carbon dioxide. Dowell is licensed to use these foams with payment of
license fees (USA and Canada only).
Binary Foam
Binary (dual-gas) foam is not subject to the Halliburton patent, as long as the carbon
dioxide content remains below 53%. The carbon dioxide phase in a binary foam is
usually held constant at 50% quality with the remainder made up of nitrogen.

2.1.2.1 Gas Behavior


The behavior of gases at given pressures and temperatures is predictable by use of
the ideal gas law and the gas deviation factor. Eq. 1 illustrates the ideal gas law.
PV = nRT (1)
Where:
P = pressure (psi)
V = volume (ft3)
n = number of moles (lbm/molecular wt of the gas)
T = temperature (°R)
 ft 3 psi 
R = universal gas constant:  10.73 T  .
 n 
Eq. 1 is true for moderate pressure (less than 500 psi) and low temperature. When
pressure and temperature are increased, the actual volume that the gas will occupy
deviates from the prediction of the ideal gas law. The gas law can be corrected by
the use of the “Z” factor (gas deviation factor or compressibility factor). Each gas
has a different value for Z at a given pressure and temperature because Z is based
on the critical pressure and critical temperature. The critical pressure and critical
temperature are distinctive properties of a specific element. For nitrogen, Z will
generally range from 0.8 to 1.9. A value of 1.0 assumes ideal behavior (no deviation
from PV = nRT). Eq. 1 can be rewritten as PV = ZnRT to correct for the volume
change.
If temperature or pressure is held constant and the other is varied, the number of
gas molecules required to fill a unit volume changes. An increase in temperature
decreases the number of molecules required to fill a unit volume when pressure is
constant. An increase in pressure increases the number of molecules required to fill
a unit volume when the temperature is held constant.
By manipulating the ideal gas law, the change in volume can be predicted.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 10 of 30

Example
Psc Vsc PV
=
Z sc Tsc ZT
thus,
PscVsc ( ZT )
V= = ft 3
Tsc Z sc P
subscript "sc" means standard conditions.
If the values for standard conditions (14.7 psi and 60°F) are substituted in the
example and Vsc is assumed to be equal to unity, then the equation becomes Eq. 2.
 P
198.6   = scf / bbl (2)
 ZT 
Where:
P = pressure (psi)
T = temperature (°R)
Z = gas deviation factor (dimensionless).
Volume factors and gas deviation factors are provided in the Nitrogen Engineering
Handbook and A Practical Companion to Reservoir Stimulation, Section P.
Obviously, an accurate value for bottomhole fracture pressure is critical because all
compressibility calculations are based on this value. Nitrogen and carbon dioxide
occupy a different amount of space for any given pressure. The volume of these
gases must be calculated at fracturing pressures. Changing the amount of gas
pumped based on surface pressures during a treatment will almost always lead to an
error in the downhole foam quality.

2.1.2.2 Gas Solubility


Nitrogen and carbon dioxide are soluble in the liquid phase (refer to Table 2.). In
water, the amount of gas lost to the solution is generally considered not significant
when calculating gas volumes. In oil, the carbon dioxide solubility is high and the
amount of gas lost to solution should be accounted for. The FracCADE* software
considers gas solubility in the foam calculations. The total volume of gas needed for
the quality calculation is the sum of the volume of the gas at fracturing conditions
and the solubility.
Gas solubilities of nitrogen and carbon dioxide in crude oil and water are provided in
A Practical Companion to Reservoir Stimulation, Section P.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 11 of 30

2.1.3 Foaming Agent Selection


The foaming agents that are typically used in foam fracturing fluids are:
• EZEFLO∗ F40 Surfactant
• Foaming Agent F52.1
• EZEFLO F78 Surfactant
• a F78/EZEFLO F75N Surfactant Mixture.
Selection of a foaming agent is usually determined by the type of foam fracturing
fluid and the foaming agent compatibility with the formation fluids. Testing of the
foaming agent with the formation fluids is strongly recommended prior to performing
the fracturing treatment in an unfamiliar area.

2.1.3.1 Material Compatibility with Foaming Agents


Materials that contaminate and inhibit the performance of the foaming agents should
be avoided. These include:
• antifoam agents or defoamer
• hydrocarbons (as additives or in the formation)
• heavy brines (including formation brines)
• alcohols
• mutual solvents.
Foaming agents can tolerate contaminants in concentrations of one to three percent
in the aqueous phase. Compatibility testing is recommended prior to using any
materials that inhibit the performance of the foaming agent.
Formation Fluids
Formation fluids may not only affect the performance of the foam but can also create
an emulsion block in the formation.
Polymer Slurries
F78 and clay stabilizers such as Clay Stabilizer L55 are not recommended additives
in a foam fracturing fluids containing HPG Polymer Slurry J876, PSG Polymer Slurry
J877, or diesel-base slurries (for example, fluid-loss additive slurries). A loss in foam
stability may occur with some field mix-waters. Laboratory testing should be
performed prior to using F78 or clay stabilizers in a foam fracturing fluid containing
J876, J877, or diesel-base slurries.
Diesel-base slurries should always be used with caution when preparing a foam
fracturing fluid. The diesel phase can be detrimental to the foam stability in some

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 12 of 30

mix-waters. Thorough prejob testing must be performed using field mix-water prior
to pumping to ensure foam stability.

2.2 Foam Rheology


The rheology of a foam is dependent on many factors, the most important being;
• foam quality
• foam texture and/or mixing
• temperature
• liquid-phase composition (polymer type and concentration).
Foam quality (discussed in Section 2.5) affects the shape and strength of the bubble
interface structure which in turn affects the viscosity of the foam. The effect of foam
quality on viscosity is illustrated in Fig. 3.
Foam texture is the average bubble size and bubble-size distribution. A foam with
smaller average bubble sizes has more interfaces formed, and therefore has a
higher viscosity. Also, the more uniform the bubble sizes, the higher the foam
viscosity. Foam texture and its dependence on controllable factors including mixing
energy is discussed in Section 2.6 and Section 3.1.

Fig. 3. The effect of foam quality on viscosity (StableFOAM fluid).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 13 of 30

2.3 Fluid-Loss Properties


Foam fracturing fluids can have two mechanisms of fluid loss in effect, two-phase
behavior of the foam and wall-building effects.

2.3.1 Two-Phase Behavior of the Foam


The most common fluid-loss-control mechanism associated with foam fracturing fluid
is the two-phase behavior of the foam. This two-phase effect helps control fluid loss
by increasing the flow resistance of the foam through the matrix of the rock. The
structure of the foam remains intact while flowing through a pipe or in the fracture.
However, the structure of the foam must deform to flow into the formation. The
deformation creates resistance and is illustrated in Fig. 4. The bubbles must deform
to leak off into the rock matrix which has small openings (smaller than the bubble
size). This deformation requires much more energy than the leakoff of a fluid with a
single-phase behavior.
The two-phase fluid-loss control mechanism is lost once the pore throat-size of the
formation exceeds the bubble size. Based on laboratory tests, this occurs at a
permeability of approximately 30 to 50 md. Above this range, foams exhibit poor
fluid-loss control.

Fig. 4. Leakoff of a foam into the rock matrix.

2.3.2 Wall-Building Effects


A foam containing a polymer can control fluid loss by filter-cake deposition on the
fracture face. This is particularly true for low-quality foams or foams containing high
concentrations of polymer. Anytime a wall-building polymer such as guar or HPG is
used, there is some degree of filter-cake deposition. Wall-building properties of foam
that contain low concentrations of polymer or high-quality foams are slow to develop.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 14 of 30

In small treatments, the treatment may be finished before the filter cake is thick
enough to have a significant effect on the fluid loss.
Methodology
Foam leakoff is fluid-phase dominated. The leakoff values for foams are lower than
the leakoff values of crosslinked water-base fluids containing diesel as a fluid-loss
additive in low permeability (<1 md) reservoirs. However, the best approach to use
when considering fluid loss in a fracture design using a foam fluid is to use the fluid-
loss coefficient measured in the laboratory as a wall-building coefficient (Cw). Unless
otherwise indicated by laboratory data, the spurt is zero and the leakoff viscosity is
that of water (1 cp).
Foam quality has little effect on Cw at low permeabilities (0.1 to 1 md). Cw decreases
as quality increases at higher permeabilities.

2.4 Conductivity Damage


Permeability of a proppant pack may be damaged by the polymer used to viscosify
the aqueous phase. Proppant-pack damage results from polymer concentration
within the proppant pack due to fluid leakoff during the treatment and the fracture
volume reduction during closure. The final postclosure polymer concentration can
easily reach 10 to 15 times the initial polymer concentration in Waterfrac (WF) fluids.
Foam is considered to be very good at minimizing proppant-pack conductivity
damage caused by postclosure polymer concentration. This is due to the low
polymer concentration in the foam. For example, a 70% quality foam containing a
liquid-phase polymer concentration of 20 lbm/1000 gal has an equivalent viscosity of
a WF fluid containing polymer at 40 lbm/1000 gal. The combination of a low polymer
concentration in the liquid-phase and less liquid-phase volume yields an effective
polymer concentration of 6 lbm/1000 gal. This foam has 85% less polymer and is
clearly a less damaging fluid. Laboratory results have shown 200% conductivity
improvements when foam is used as a fracturing fluid instead of a WF fluid.
Proppant-pack damage can be determined using the following procedure.
1. Determine the average proppant concentration (lbm proppant added per gal) for
the fracturing treatment.
2. Determine the polymer concentration factor from Fig. 5. The proppant-pack
porosity is dependent on the closure stress. Table 3 provides a guide for
determining the porosity of sand and intermediate-strength proppant.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 15 of 30

Table 3. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant


Porosity (%) Closure Stress Sand (psi) Closure Stress ISP (psi)
37.0 1000 4000
33.5 3000 6000
30.0 5000 8000

Fig. 5. Dimensionless polymer concentration factor.


3. Determine the polymer concentration in the fracture using Eq. 3.
Ce = C paq (1 − Γ )C pol (3)

Where:
Ce = effective polymer concentration (lbm/1000 gal)
CPaq = aqueous phase average polymer concentration (lbm/1000 gal)
Γ = average bottomhole foam quality (decimal)
Cpol = polymer concentration factor (dimensionless).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 16 of 30

4. Determine the retained permeability from Fig. 6.


This retained permeability is based on the damage caused by the polymer and not
by other factors such as embedment and fines.

Fig. 6. Polymer concentration versus proppant-pack retained permeability.

2.5 Foam Quality


Foam quality is defined as the volumetric fraction (expressed as a percentage) of
gas. A 70% quality foam contains 70% gas and 30% vol by liquid. If the foam
quality is low, the gas bubbles can freely move within the liquid. At higher foam
qualities, the bubble concentration is sufficient to hinder segregation, a process
similar to the hindered settling effects of proppants.
Fig. 7 illustrates the bubble arrangements for various foam-quality ranges.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 17 of 30

Fig. 7. Bubble arrangements for various foam-quality ranges.


Less than 52% Quality
At less than 52% quality, the spherical bubbles are free to segregate and move.
This configuration is considered unstable. The limiting geometry is where bubbles of
equal size occupy cubes of unit volume dimension. The gas sphere within the cube
occupies a volume equal to π/6 of the unit volume. This results in foam quality equal
to 52% which assumes equal spherical bubble-size distribution. Fluids in which the
gas phase is less than 52% are classified as energized fluids.
From 53% to 74% Quality
From 53% to 74% quality, the spherical bubbles contact each other and are no
longer free to move.
From 74% to 96% Quality
From 74% to 96% quality, the bubble sides flatten and there is a more rigid structure.
Greater than 96% Quality
At greater than 96% quality, a gas-outside-phase and liquid-inside-phase are formed
resulting in a mist. Fluids in which the gas phase is greater than 96% are classified
as atomized fluids.
Experiments have shown that as polymer loading increases, foam stability increases
because of the viscosity increase in the liquid phase. The minimum quantity of gas
required to make a stable foam decreases. For example, a liquid phase containing

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 18 of 30

hydroxypropylguar (HPG) at 40 lbm/1000 gal may be stable at a foam quality as low


as 40%.
The structure formed by the network of bubbles is strengthened at higher foam
qualities. This is caused by the liquid phase being stretched further over the gas
phase. This effect reduces the free liquid associated with the foam and makes the
bubble interfaces more rigid. The spherical shape of the bubble will change to a
polyhedral shape where the interfaces are much flatter as shown in. High foam
qualities may be desirable because of the following.
• improved proppant transport
• higher viscosity
• better fluid-loss control
• less fracture conductivity damage because of decreased quantities of liquid.
The most significant drawback to high foam qualities is the difficulty in achieving high
proppant concentrations. Fig. 8 illustrates the effect of increasing the foam quality
on the proppant concentration. A slurry concentration of 20 lbm proppant added
(PPA) at the blender would yield a concentration of 2 PPA in a 90% quality foam.

Fig. 8. Proppant concentration limits in foam fluids.


Foams made with crosslinked fluids as the liquid phase are usually thought of as
being stable at any quality, because the high viscosity associated with the
crosslinked structure prevents the gas bubbles from segregating from the liquid.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 19 of 30

However, there may be limits on the maximum quality a crosslinked foam may be
stable. Depending on the crosslinker type and concentration, the crosslinked
structure of the liquid phase may be so rigid that it resists being disrupted by a large
bubble concentration. If this occurs, any additional gas will simply separate from the
foam. A 70% quality foam appears to be the maximum for most crosslinked foams.
However, there is no advantage gained by crosslinking a foam unless the quality is
less than 52%.

2.6 Foam Texture


Foam texture is the average bubble size and bubble-size distribution. Foam texture
is an important factor in determining rheology and fluid-loss properties. In general, a
foam containing small bubbles is more viscous than the same quality foam
containing larger bubbles. Likewise, a foam containing a narrow range of bubble
sizes is more viscous than a foam containing a wide range of bubble sizes. These
generalities relate back to the kind and strength of the bubble interface structure that
is created within the foam. Normally, these factors are not controllable in the field
except by the foaming agent type and concentration or the viscosity of the liquid
phase. For example, a high-quality crosslinked foam may exhibit a very coarse
average bubble-size. While this type of a fluid may be desirable, it will behave more
like a crosslinked fluid instead of a foamed fluid because of a lack of foam structure.
If the foam fluid is pumped in turbulent flow, sufficient mixing turbulence is available
to create a good foam texture. Turbulent flow may not be achieved at low pump
rates or with high-viscosity fluids. The gas phase must be thoroughly dispersed in
SuperFOAM fluids before crosslinker is added. Pumping the crosslinker
downstream of the gas will allow foaming to occur when the liquid-phase viscosity is
low and will ensure good foam texture.

2.7 Proppant Compensation


Hydraulic fracturing fluids that contain proppant are thought of as slurries. Slurries
are composed of a fluid and a solid. As more proppant is added to the slurry, the
quantity of liquid required decreases due to the additional volume taken up by the
proppant. The proppant volume will displace an equivalent volume of fluid. In foams,
the proppant can act as part of the internal phase, making the foam behave as a
higher quality foam.
The proppant volume can have a considerable effect on the foam quality. One of the
following methods is generally used when compensating for the proppant volume.
• No compensation
• Constant bottomhole quality
• Decreasing bottomhole quality
• Constant internal phase (CIP).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 20 of 30

In all cases but no compensation, the purpose of the compensation is to allow


increasing proppant concentrations while maintaining adequate but not excessive
viscosity. There is no apparent technical rationalization of preferring any of the
above.
Fig. 9 illustrates the effect of proppant compensation methods on bottomhole foam
quality for a 70% quality foam and a proppant specific gravity of 2.65.

Fig. 9. The effect of proppant compensation methods on bottomhole foam quality.

2.7.1 No Proppant Compensation


The only advantage of no compensation is that it is operationally simple. As the
proppant concentration at the blender increases, the total slurry rate increases.
These increases may become so large that the limitations of the pumping equipment
will be exceeded. However, in jobs containing stages with small volumes of
proppant, it may be operationally more simple to not compensate for the proppant in
some consecutive stages and then make a foam-quality correction before a problem
arises.
The actual foam quality for foam having no compensation for the proppant can be
determined with Eq. 4.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 21 of 30

Γd − x
Γa = (4)
1− x

Where: X=
( 1 − Γd )C pfoam
8.34 SG p

and
Γa = actual foam quality (decimal)
Γd = designed foam quality (decimal)
Cpfoam = proppant concentration in the foam (PPA)
SGp = specific gravity of proppant (dimensionless).

2.7.2 Constant Bottomhole Quality


Constant bottomhole quality (constant gas-to-liquid volume ratio) is usually the
chosen method. When a job is designed, the constant bottomhole quality in the
fracture is often assumed. This quality is accomplished by one of three methods.
• Gas rate adjustment
• Liquid rate adjustment
• Gas and liquid rate adjustment.
Adjusting the pump rate of the gas phase involves reducing the gas pump rate as the
liquid volume decreases and the proppant volume increases. The blender discharge
rate remains constant. This type of adjustment will lower the total pump rate as the
job progresses.
Adjusting the pump rate of the liquid phase involves increasing the liquid rate to
compensate for the liquid volume displaced by the proppant volume. The gas rate
remains constant. This type of adjustment will increase the total pump rate as the
job progresses.
Adjusting the pump rate of the gas or liquid (described above) may vary the total flow
rate by as much as 20%. The total flow rate and foam quality will remain constant by
simultaneously adjusting the gas and liquid rate.

2.7.3 Decreasing Bottomhole Quality


The bottomhole quality can be reduced to accommodate higher proppant
concentration For example:
1. Pad, 70% quality
2. 2 PPA (wellhead), 70% quality
3. 4 PPA (wellhead), 68% quality
4. 6 PPA (wellhead), 65% quality.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 22 of 30

2.7.4 Constant Internal Phase


In the constant internal phase method, equal volumes of gas are removed as
proppant is added, resulting in a constant ratio of the internal phase volume to the
total volume. Therefore, as the proppant concentration increases and the gas
volume decreases, the foam quality decreases.
For constant internal phase:
Gas Volume + Pr oppant Volume
Foam Quality =
Gas Volume + Proppant Volume + Liquid Volume
The constant internal phase method for proppant compensation is patented by
Halliburton Services. Using this method with either nitrogen or carbon dioxide foams
requires payment of a license fee. The constant internal phase is a method
developed to approximate a constant viscosity foam or emulsion regardless of the
proppant concentration. The procedure, as outlined in Halliburton's patent, does not
result in constant fluid viscosity nor does it imply the most efficient proppant
transport. Maintaining a constant viscosity is usually not a problem in most foam
fracturing treatments. The change in viscosity is minor because most treatments
have relatively low proppant concentrations (less than 10 PPA) and foam bubbles do
not act as particles (some emulsions have liquids that act as particles). Other
parameters such as the temperature and gas volume ratio can affect the foam
viscosity more than the proppant concentration.
Dowell personnel refrain from designing a foam fracturing treatment using the
concept of constant internal phase as patented. One of the other methods will be
adequate for any treatment design.
The actual foam quality for foam having a constant internal phase can be determined
with Eq. 5.
Γa = Γd − x (5)

Where: X=
( 1 − Γd )C pfoam
8.34 SG p

and
Γa = actual foam quality (decimal)
Γd = designed foam quality (decimal)
CPfoam = proppant concentration in the foam (PPA)
SGp = specific gravity of proppant (dimensionless).

2.8 Friction Pressure


Foams, like crosslinked fluids, are difficult to scale up from laboratory to field
conditions. This is particularly evident with friction pressures. The unique

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 23 of 30

rheological behavior and the difficulty encountered matching field conditions


contribute to the problem. Therefore, the friction pressure data provided in the
Fracturing Materials Manual  Fluids are only approximate.

2.9 Yield Stress


A yield stress represents the amount of force that is applied to a fluid before the fluid
yields or begins to flow. Many different types of foams exhibit a yield stress. For
example, suds in a sink can support small lightweight objects. In this case, the force
applied by the object is not large enough to overcome the yield stress of the foam.
Most foam fracturing fluids exhibit a yield stress. However, the behavior of the
flowing foam in a fracture is largely unaffected and can effectively be modeled as a
power-law fluid (no yield stress). The yield stress becomes most important at very
low shear rates (less than 10 sec-1). Therefore, the proppant suspension properties
of a foam are enhanced. Because of this yield stress, foams have excellent
proppant suspension characteristics, similar to a crosslinked fluid.

2.10 Limitations of Application


The inability to pump high proppant concentrations in high-quality foams is a
limitation in foam fracturing treatments. Using a low-quality foam may not be
desirable because of formation or fracture conductivity damage. Varying the foam
quality throughout the treatment is a good compromise approach. For example, a
70% quality foam can be used for the pad and low-proppant concentration stages
(6 PPA and less), followed by a 50% quality foam (10 PPA and less). High proppant
concentrations in low-quality StableFOAM fluids (12 PPA in 30% Quality
StableFOAM) are not recommended. The foam structure will not properly transport
the proppant in the fracture. SuperFOAM must be used to obtain a high proppant
concentration in a low-quality foam.

2.11 Job Design


The Foam Schedule module is in the FracCADE software. The module allows a
variety of treatment combinations to be simulated until the optimum design is
obtained. The FracCADE Users Manual and FracCADE Release Notes provide
additional foam fracturing job design information.

2.12 Calculations

2.12.1 Pressures and Rates


1. Determine the reservoir fracturing pressure (Eq. 6).
piw = p fg Dp (6)

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 24 of 30

2. Determine the hydrostatic pressure of the liquid phase (Eq. 7).


p Lh = phg SG L δ L D p (7)

3. Determine the hydrostatic pressure of the gas phase (Eq. 8). Use the tables
provided in Section 315 in the Nitrogen Engineering Handbook to determine the
wellhead pressure.
pgh = δ g ( piw − pwh ) (8)

4. Determine the hydrostatic pressure of the foam fluid without proppant (Eq. 9).
p fh = p Lh + pgh (9)

5. Determine the foam friction-pressure (without proppant) using the friction


pressure data provided in the Fracturing Materials Manual  Fluids.
6. Determine the flow rate through each perforation by dividing the foam pump-rate
by the number of perforations.
7. Determine the perforation friction pressure (ppf) using Fig. 10.
8. Determine the surface pumping pressure at the wellhead (Eq. 10).
pw = piw − p fh + ptf + p pf (10)

Where:
Dp = true vertical depth to mid-perforation (ft)
pfg = fracture gradient (psi/ft)
pfh = hydrostatic pressure of the foam (psi)
pgh = hydrostatic pressure of the gas phase (psi)
phg = hydrostatic pressure gradient (psi/ft)
piw = reservoir fracturing pressure (psi)
pLh = hydrostatic pressure of the liquid phase (psi)
ppf = perforation friction pressure (psi)
ptf = tubular friction pressure (psi)
pw = surface fracturing pressure (psi)
pwh = wellhead pressure (psi)
SGL = specific gravity or the liquid phase
δL = liquid ratio (decimal)
δg = gas ratio (decimal).

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 25 of 30

Fig. 10. Friction through perforations.

2.12.2 Equipment Requirements


1. Determine the liquid-phase pump rate (Eq. 11).
q L = ( qδ L ) − q f (11)

2. Determine the horsepower requirement for the liquid phase (Eq. 12).
HHP = 0.0245 pw q L (12)
3. Determine the gas-phase pump rate (Eq. 13).
q g = qδ g (13)

4. If the gas phase is nitrogen, determine the nitrogen pump rate in scf/min by
dividing the nitrogen pump rate found in step 3 by the nitrogen volume factor.
Nitrogen volume factors are provided in Section 3 of the Nitrogen Engineering
Handbook.
5. If the gas phase is carbon dioxide, use Eq. 11 and Eq. 12 to determine the pump
rate and horsepower requirements.
6. Determine the number of nitrogen pumpers required by dividing the nitrogen
pump rate (scf/min) by rate per pumper.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 26 of 30

Where:
HHP = hydraulic horsepower
pw = surface fracturing pressure (psi)
q =pump rate (bbl/min)
qf = foamer pump rate (bbl/min)
qg = gas-phase pump rate (bbl/min)
qL = liquid-phase pump rate (bbl/min)
δL = liquid ratio (decimal)
δg = gas ratio (decimal).

2.12.3 Material Requirements


1. Determine the total liquid-phase volume (gal) by multiplying the total foam
volume (gal) by the liquid ratio. Adjust for proppant volume if required.
2. If the gas phase is nitrogen, determine the total nitrogen volume (SCF) by
multiplying together the total foam volume (gal), the gas ratio, the nitrogen
volume factor (scf/bbl) and 0.0238. Adjust for proppant volume if required.
3. If the gas phase is carbon dioxide, determine the total carbon dioxide volume
(gal) by multiplying the total foam volume (gal) by the gas ratio. Use the
conversion factors for carbon dioxide provided in Section 600 of the Nitrogen
Manual to convert gallons to tons. Adjust for proppant volume if required.
4. Determine the total foamer volume (gal) by multiplying together the foamer
concentration (gal/1000 gal), the total liquid-phase volume (gal) and 0.001.
5. Determine the material requirements for additional additives.

3 Execution
Designing a foam fracturing job is similar to any other fracturing job with the following
exceptions.
• equipment and location limitations
• foam generation
• proppant compensation methods.
Executing a foam fracturing job requires a significant amount of equipment and
pump-schedule coordination. Jobs should be well planned in advance to assure all
pumping and storage requirements can be met. In some applications, foam
parameters such as quality may be limited because of available equipment or size of
the location. Fig. 11 is a schematic of a typical foam fracturing treatment.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 27 of 30

Fig. 11. Schematic of foam fracturing treatment.


Pressure Measurement
For pressure analysis during a fracturing treatment, the bottomhole pressure can be
calculated from the surface pressure as long as the fluid density is constant and the
bottomhole pressure is greater than the hydrostatic pressure. The main problem
with using the treating pressure for analysis is that the friction pressure makes the
Nolte-Smith plot less accurate and can indicate erroneous trends. In the overall
analysis, the Nolte-Smith plot is very valuable if accurate bottomhole pressure and
closure pressure are used.
A good method for measuring bottomhole pressure is with a “live” annulus or a
“dead-string tubing” and a homogeneous fluid. This eliminates friction pressure
calculations. With a known hydrostatic pressure, bottomhole pressure can be
accurately calculated. The density of the static column of fluid must be known
(circulate the well and check the specific gravity of the fluid prior to injection). The
fluid must not contain any trapped gas. This method is generally adequate for wells
with a bottomhole static temperature less than 250°F (121°C) and a depth less than
10,000 ft.
Direct measurement of downhole treating pressure is not generally used with
fracturing treatments using proppant. The sensor and wireline can become stuck in
the event of a screenout.
Proppant Concentration
The inability to mix and meter high proppant concentrations in high-quality foams is a
limitation in foam fracturing treatments because of the dilution effects of the gas

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 28 of 30

phase. A hydrocyclone separator (sand concentrator) was once used in the high-
pressure discharge line to counteract the dilution effects and increase the proppant
concentration. Introduction of the POD* programmable optimum density blender has
obsoleted the sand concentrator by eliminating the operational complexities
associated with the concentrator. The proppant concentration can be precisely
controlled up to 22 PPA (sand) and 32 PPA (high-strength proppant) with a POD
blender.
Flush Fluids
Do not use foam fluid to flush a foam fracturing stage containing proppant. The
foam fluid will heat-up and expand while the fracture is closing and may cause the
proppant to be displaced away from the wellbore.

3.1 Foam Generation


Foam generation is usually not a problem under typical fracturing flow conditions. If
the fluid is pumped in turbulent flow, sufficient mixing turbulence is usually available
to create a good foam texture. Turbulent flow may not be achieved at low pump
rates or with high-viscosity fluids.
Flow characteristics of foamed fluids may be determined by reference to the
respective friction pressure charts. A foam generator is required if the friction
pressure charts indicate that the foamed fluid will be in laminar flow or anytime the
total pump rate is less than 10 bbl/min. Fig. 12 illustrates typical laminar and
turbulent flow areas of foamed fluids. A foam generator is recommended for all foam
fracturing treatments, especially those using carbon dioxide as the gas phase.
A foam generator is a frac cross containing a disperser through which the gas phase
is pumped. The disperser comes with various hole sizes and arrangements.
Typically, there are 16 holes for nitrogen and 48 holes for carbon dioxide. The
number and size of holes will be determined by the treatment pump rate. The
disperser also contains a burst disc. A foam generator is illustrated in Fig. 13.
Additional foam generator information is provided in Standard 9 in the Dowell Safety
and Loss Prevention Manual.
Effects of Surfactant  Foam that contains an insufficient quantity of surfactant or
that has contaminated surfactant may have a poor foam texture. Friction pressure
that is lower than expected is often a strong indicator of a foam texture or mixing
problem.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Foam Fracturing
Page 29 of 30

Fig. 12. Laminar and turbulent flow areas of foamed fluids.

Fig. 13. Foam generator.

DOWELL CONFIDENTIAL
Section 700.2
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Foam Fracturing Dowell
Page 30 of 30

3.2 Material Balance


Planning the material balance of a foam fracturing treatment is one of the most
critical steps toward precise execution. The supervisor in charge of pumping
operations must know exactly how much material is needed, how much has already
been pumped and how much remains on location. A pumping schedule (plan)
should be organized prior to the fracturing treatment, and will allow tracking
throughout the treatment. During the treatment, adjustments to the pumping
schedule can be made for any deviation from the plan.
A pumping schedule is generated by the foam schedule module in the FracCADE
software. A pumping schedule containing the following columns may be hand-
generated.
• Step number
• Pump time (min)
• Liquid-phase pump rate (bbl/min)
• Liquid-phase volume (bbl)
• Liquid-phase cumulative volume (bbl)
• Foamer pump rate (bbl/min)
• Foamer volume (bbl)
• Foamer cumulative volume (bbl)
• Gas-phase pump rate (scf/min or bbl/min)
• Gas-phase volume (scf or bbl)
• Gas-phase cumulative volume (scf or bbl)
• Proppant concentration at the blender (PPA)
• Proppant volume (lbm)
• Proppant cumulative volume (lbm)
• Proppant concentration in the foam (PPA)
• Foam quality (%).

DOWELL CONFIDENTIAL
Section 700.3
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
RampGEL Service
Page 1 of 5

RampGEL SERVICE

1 Introductory Summary............................................................................................................. 1
1.1 Application ........................................................................................................................... 2

2 Design ....................................................................................................................................... 3
2.1 Pad Fluid.............................................................................................................................. 4
2.2 Fracturing Fluid Transporting Proppant ............................................................................... 4
2.3 Computer-Aided Design....................................................................................................... 4

3 EXECUTION .............................................................................................................................. 4
3.1 Batch Mixing ........................................................................................................................ 4
3.2 Continuous Mixing ............................................................................................................... 4
FIGURES
Fig. 1. Polymer concentration design scheme for the RampGEL software.................................. 5

1 Introductory Summary
The RampGEL* Service is a technique that optimizes the fracturing fluid polymer
concentration. The RampGEL service will provide the minimum amount of polymer
necessary (per stage) to maintain adequate fracturing fluid viscosity and fluid-loss
control for fracture propagation and proppant transport. The benefits of this
technique are
• increased treatment efficiency (no excess polymer)
• improved well performance (lower postclosure polymer concentration resulting in
lower proppant-pack conductivity damage)
• lower treatment cost.
Achieving the desired fracture length, height, width, and conductivity in the most cost
effective manner is one of the major challenges in fracturing treatment design.
Fracturing fluid viscosity and leakoff characteristics strongly govern the fracture
propagation behavior and the distribution and placement of proppants. The addition
of polymer, stabilizers, and fluid-loss additives to a fracturing fluid increases
proppant-transport properties (by increasing and maintaining viscosity) and

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.3
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
RampGEL Service Dowell
Page 2 of 5

decreases fluid loss. If the polymer concentration is insufficient, the treatment may
terminate early because of fluid depletion or inadequate proppant transport.
However, an excessive amount of polymer may cause unnecessary proppant-pack
conductivity damage and may increase the potential for undesired vertical height
growth. Excess polymer increases treatment costs and lowers the net present value
(NPV). Thus, it becomes important to design a treatment utilizing an optimized
polymer concentration.
Rather than relying on some empirical design guidelines, sound bases have been
developed to schedule polymer concentration. The optimal polymer concentration is
selected based on:
• fluid rheology
• fluid efficiency
• proppant type
• proppant concentration
• reservoir temperature
• fluid temperature
• fluid shear-exposure history.
To ensure proper fluid properties in the fracture, the RampGEL software considers
two important functions of polymers; fluid-loss control and proppant transport.
Separate criteria have been adopted to design polymer concentration for the pad
volume and slurry volume because each serves a different purpose. The polymer
concentration for the pad volume is designed primarily for fluid-loss control. To
maximize treatment efficiency, the optimum polymer concentration is determined by
maximizing the ratio of fracture-volume to polymer-weight (cost). The polymer
concentration for the slurry volume is designed by using the minimum viscosity
required for proppant transport. This design will simultaneously ensure adequate
viscosity for proppant transport and use the pad fluid to more effectively control fluid
leakoff.

1.1 Application
The RampGEL software is used to determine the optimal polymer concentration
schedule at given treatment conditions. It supports the Perkins-Kern-Nordgren
(PKN), Khristianovic-Geertsma de Klerk (KGD) and radial fracture models through
the Fracture Geometry Sensitivity (FGS) module in the FracCADE* software. The
polymer concentration can be scheduled based on either fracture penetration or total
fluid volume (time). A desired fracture penetration may be obtained from the
FracNPV* software. The RampGEL software also has the option, for batch mixed
treatments, to specify a minimum fluid volume (tank size) for each pumping stage.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.3
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
RampGEL Service
Page 3 of 5

Currently, application of the RampGEL software is limited to the YF*500HT,


YF600LT, and YF600HT series fracturing fluids.

2 Design
A schematic representation of the RampGEL design process is shown in Fig. 1. A
treatment can be designed based on either fracture penetration or total fluid volume
(or total pumping time if the pump rate is constant). This is a design feature
supported by FGS.
Polymer concentration design scheme for the RampGEL software. The three main
calculating units in the RampGEL software are:
• the polymer-concentration design module
• the fluid-tracking module
• the polymer-concentration schedule module.
The polymer-concentration design module calculates the polymer efficiency as a
function of polymer concentration using FGS. For each polymer concentration, FGS
is used to simulate fracture geometries and treatment parameters. The optimum
polymer concentration for the pad fluid is then determined by maximizing the
polymer efficiency. The polymer efficiency is defined as

Fracture Volume
Polymer Efficiency =
Total Polymer Used

Fracture Volume
=
TotalFluid Pumped x Polymer Loading

n
= 1000
Cpolavg

Where:
n = fluid efficiency (dimensionless)
Cpolavg = average polymer concentration (lbm/1000 gal)
1000 = conversion factor.
Once the optimum polymer concentration is determined, the corresponding fluid
efficiency, pad fraction, and fracture propagation parameter are used by the fluid-
tracking module to calculate fluid exposure time in relation to the pump time.
After the exposure-time profile is obtained, the polymer-concentration schedule
module is used to select the polymer concentration for the pad fluid and the slurry
according to specified conditions.

DOWELL CONFIDENTIAL
Section 700.3
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
RampGEL Service Dowell
Page 4 of 5

The pad fluid will always be designed to have at least the viscosity of the first stage
of proppant to assure continuation of slurry. This implies that the pad polymer
concentration is always equal to or greater than the polymer concentration used in
the first stage of proppant.

2.1 Pad Fluid


If a uniform polymer concentration is desired for the pad fluid, the optimum polymer
concentration is used as the polymer concentration for all fluids in the pad. If a
ramped polymer concentration (with respect to time and temperature) is desired then
the polymer concentration is calculated.

2.2 Fracturing Fluid Transporting Proppant


The polymer concentration for the slurry is selected as the minimum concentration
required to maintain a constant proppant settling rate (or a uniform viscosity) at the
end of pumping. The end viscosity is calculated if a settling rate is specified. For a
given temperature and exposure time, the fluid viscosity is a function of polymer
concentration only. Once the end viscosity is known, the corresponding polymer
concentration may be found directly from the fluid rheology database. The proppant
fall distance is not used as a polymer concentration design criterion.

2.3 Computer-Aided Design


The RampGEL design is incorporated into the Inverse module in the FracCADE
software. Additional design information is provided in the FracCADE/VAX User
Manual and FracCADE V2.6 Release Notes.

3 EXECUTION

3.1 Batch Mixing


The number of different polymer concentrations and volumes is dependent on the
quantity of tanks and tank volumes.

3.2 Continuous Mixing


Ramping the polymer concentration is accomplished with the PCM* precision
continuous mixer. The PCM is a pumping and blending system that allows
continuous mixing of fracturing fluid. Water-base fracturing fluids with equivalent
polymer concentrations of 10 to 60 lbm/1000 gal may be continuously mixed and
discharged at a maximum of 70 bbl/min. Additional information is provided in the
PCM Operator's Manual.
A slurried polymer metering unit and hydration tank may also be used for
continuous-mix applications.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.3
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
RampGEL Service
Page 5 of 5

Fig. 1. Polymer concentration design scheme for the RampGEL software.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 1 of 20

CleanFRAC SERVICE

1 Introductory Summary............................................................................................................. 2
1.1 Applications.......................................................................................................................... 2

2 Design ....................................................................................................................................... 2
2.1 Well Candidate Selection ..................................................................................................... 3
2.2 EB-Clean J475 Breaker ....................................................................................................... 3
2.2.1 Mechanism of Breaker Release.................................................................................. 5
2.3 Breaker Selection................................................................................................................. 6
2.4 Job Design ........................................................................................................................... 7
2.4.1 CleanFRAC Service Design Methods......................................................................... 7
2.4.2 Summary of Software Features for the CleanFRAC Service Design.......................... 7
2.4.3 Design Procedure Using a Manual Calculation .......................................................... 8
2.4.4 Design Procedure Using the FracCADE Software Exclusively or in Combination
With a Manual Calculation........................................................................................ 11

3 Execution................................................................................................................................ 12
3.1 Metering the J475 Using the Auger-Type, Dry-Additive Feeder ........................................ 12
3.1.1 Procedure for Scheduling the J475 Using the Auger-Type Dry-Additive Feeder ..... 13
3.2 Metering the J218 as an Aqueous Solution ....................................................................... 14
3.2.1 Procedure for Calculating Mixing and Metering Parameters for J218 Solutions ...... 14

4 Examples ................................................................................................................................ 16
4.1 Example Design Using a Manual Calculation .................................................................... 16
4.2 Example Procedure for Scheduling the J475 Using the Auger-Type
Dry-Additive Feeder........................................................................................................... 18
4.3 Example Procedure for Calculating Mixing and Metering Parameters for
J218 Solutions ................................................................................................................... 19
FIGURES
Fig. 1. Effect of breaker concentration on retained viscosity for a borate-crosslinked guar
fluid at 160°F (71°C).......................................................................................................... 4
Fig. 2. Effect of breaker concentration on retained proppant-pack permeability.......................... 5
Fig. 3. Effects of proppant concentration and porosity on postclosure polymer concentration. ... 9

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 2 of 20

Fig. 4. Influence of ammonium persulfate breaker on retained proppant-pack permeability


for linear fracturing fluids. ..................................................................................................9
Fig. 5. Influence of ammonium persulfate breaker on retained proppant-pack permeability
for borate-crosslinked fracturing fluids. ...........................................................................10
Fig. 6. Influence of ammonium persulfate breaker on retained proppant-pack permeability
for organometallic-crosslinked fracturing fluids. ..............................................................10
TABLES
Table 1. J475 Release Levels (no closure stress)........................................................................6
Table 2. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant ..........................8
Table 3. J475 Metering Rates Auger-Type, Dry-Additive Feeder...............................................12

1 Introductory Summary
The CleanFRAC* Service is a technique that optimizes fracture conductivity by
reducing proppant-pack damage caused by highly concentrated polymer in fracturing
fluids containing guar, hydroxypropylguar (HPG) and hydroxyethylcellulose (HEC)
gelling agents. It offers an approach to quantify proppant-pack damage and a
means to reduce the damage by the use of Dowell breaker technology.

1.1 Applications
The CleanFRAC Service is most effective in stimulation services using the following
Waterfrac (WF) and Widefrac (YF*) series of fluids  WF100, WF200, YF100,
YF100D, YF200, YF200D, YF300LPH, YF400LPH, YF500HT, YF600LT and
YF600HT. The CleanFRAC Service is also effective in foam fracturing fluids.

2 Design
The CleanFRAC Service is designed to increase fracture conductivity in oil and gas
wells that would otherwise be considered conductivity limited because of the
damaging effects of water-base fracturing fluids. Polymer that is concentrated within
the proppant pack due to fluid leakoff and volume reduction during fracture closure is
the primary cause of proppant-pack damage. The severity of damage increases as
the polymer concentration increases, and is strongly dependent on fluid type,
crosslinker type and fluid efficiency. The borate-crosslinked fluids are less damaging
than the organometallic-crosslinked fluids at temperatures less than 180°F (82°C).
Breakers will reduce the severity of proppant-pack damage caused by concentrated
polymers. The amount of retained proppant-pack permeability which can be
achieved is directly related to the breaker concentration  increasing the breaker
concentration will increase the fracture conductivity.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 3 of 20

Optimum fracture conductivity, and therefore higher production rates, can be


achieved by using the CleanFRAC Service. More rapid well cleanup should also be
experienced which will result in a quicker payout and lower swabbing cost. The
CleanFRAC Service also reduces polymer damage at the fracture tip resulting in
extended fracture length.

2.1 Well Candidate Selection


A well with a fracture design having low dimensionless fracture conductivity (CfD) and
a high postclosure polymer concentration is the best candidate for the CleanFRAC
Service. High postclosure polymer concentrations produce lower fracture
conductivities resulting in lower production. The amount of conductivity damage from
water-base fracturing fluids increases with an increasing postclosure polymer
concentration, and is strongly dependent on the type of fluid used.
CfD and postclosure polymer concentrations should be used as guidelines to
determine conductivity-limited fracturing treatments. Factors that can adversely
affect CfD and/or postclosure polymer concentration are
• low fluid efficiencies and/or large pad volumes
• low proppant concentration
• moderate to high formation permeabilities
• long fracture half-lengths
• small fracture widths
• initial polymer concentrations greater than 25
• lbm/1000 gal.
A candidate for the CleanFRAC Service can also be a well in which the offset wells
have experienced lower than expected production.

2.2 EB-Clean J475 Breaker


EB-Clean* J475 Breaker is a 20/40 mesh material produced by coating
(encapsulating) an ammonium persulfate breaker material (J218) with a water-
resistant barrier. Encapsulation of the breaker greatly reduces fracturing fluid
exposure to the breaker, and enables the use of high concentrations of breaker that
without coating, would rapidly reduce the fluid viscosity. Unlike breakers which
dissolve in the fracturing fluid, for example, Breaker J218 and Breaker J134, J475
cannot leak off and be lost to the formation. It remains in the fracture where it is
needed to degrade concentrated polymer.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 4 of 20

Fig. 1 illustrates the effectiveness of the protective coating in a borate-crosslinked


guar fluid at 160°F (71°C). Note that for this fluid and temperature, J475 can be
added at a concentration of 2 lbm/1000 gal with minimal viscosity reduction, or at
4 lbm/1000 gal with an approximate 65% loss of viscosity in one hour. A minimum of
10% of the fluid viscosity would have to be maintained for effective proppant
placement for the fluid in. This would permit a maximum Breaker J218 concentration
of less than 1 lbm/1000 gal or a maximum J475 concentration of 8 lbm/1000 gal to
be used for a pumping time of one hour.

Fig. 1. Effect of breaker concentration on retained viscosity for a borate-crosslinked


guar fluid at 160°F (71°C).
Fig. 2 illustrates the effect of breaker concentration on retained proppant-pack
permeability. Note for this example that the retained proppant-pack permeability is
10% when the J218 concentration is 0.5 lbm/1000 gal and is 43% when the J475
concentration is 8 lbm/1000 gal.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 5 of 20

Fig. 2. Effect of breaker concentration on retained proppant-pack permeability.

2.2.1 Mechanism of Breaker Release


Release of the breaker from the capsule occurs in two ways.
• After completion of the fracturing treatment, the hydraulic pressure within the
fracture dissipates (due to fluid leakoff) and the fracture closes, creating high
point-to-point stresses on the proppant and J475. Faults occur in the coating that
allow water to penetrate and solubilize the ammonium persulfate. The
ammonium persulfate is then released into the concentrated fluid within the
fracture.
• Some penetration of water occurs even with highly water resistant coatings. After
sufficient exposure to water, encapsulated particles that may not be in a highly
stressed state may have sufficient water penetration to release the breaker.
Although the protective coating on the J475 is very durable, low levels of active
breaker are released during pumping. This is due to small imperfections in the
coating and damage caused by pumping equipment. Typical release levels at
various times and temperatures are shown in Table 1.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 6 of 20

Table 1. J475 Release Levels (no closure stress)


Temperature (°°F [°°C]) Time (hr) J475 Release Level (Total wt (%))
125 [52] 1 3
125 [52] 2 3
125 [52] 3 3
150 [66] 1 5
150 [66] 2 5
150 [66] 3 5
175 [79] 1 5
175 [79] 2 6
175 [79] 3 7
175 [79] 4 9
As an example, J475 may release up to 5% by wt J218 after pumping through
surface equipment and being exposed to 150°F (66°C) for one hour. Therefore,
addition of 5 lbm of J475/1000 gal fracturing fluid would result in the release of 0.25
lbm of J218/1000 gal to the fracturing fluid after one hour at 150°F (66°C).

2.3 Breaker Selection


The following guidelines are used for breaker selection.
• At temperatures less than 125°F (52°C), a combination of J218 and J318 and is
recommended.
• At temperatures between 125°F (52°C) and 170°F (77°C), a combination of J475
and J218 is recommended.
• At temperatures greater than 170°F (77°C), J475 alone is recommended. Above
180°F (82°C), J475 may rapidly deactivate.
J475 should be used in combination with J218 whenever possible, for two reasons.
• For cost purposes. It would be unreasonable to use only the more costly J475
knowing that a portion of the total breaker concentration could have been
supplemented by using the less expensive J218.
• By using a J475/J218 combination, viscosity reduction of the fracturing fluid
before closure is ensured, which reduces the shut-in and closure time.
Caution should be exercised when adding J218. Adding J218 at too high a
concentration may cause the fracturing fluid to completely break before the fracture
closure. This will cause the proppant to settle to the bottom of the fracture before

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 7 of 20

closure, and thus risk the possibility of not having a conductive fracture over the
productive zone.

2.4 Job Design


Designing an optimum breaker concentration can be performed using one of three
methods. The method chosen will primarily depend on the fracturing fluid used. A
discussion on the three methods and design procedures using the CleanFRAC
Service are provided in this section. Example breaker designs are provided in
Section 4.

2.4.1 CleanFRAC Service Design Methods


The optimum breaker schedule can be designed by using one of the following three
methods.
• Manual calculation
• FracCADE* software in combination with a manual calculation
• FracCADE software exclusively.
The simplest and preferred design method is to use the FracCADE software
exclusively to determine optimum breaker concentrations. This is the recommended
procedure and should be used whenever possible. The breaker correlations in the
FracCADE software are limited to borate-crosslinked fluids (YF100 and YF200 series
fluids) at the present time.
The design method using a combination of the FracCADE software and a manual
calculation is recommended for fluids which do not have the breaker and fluid
retained factor correlations incorporated into the FracCADE software, for example,
linear, titanate-crosslinked and zirconate-crosslinked fluids.
A manual calculation can be performed to estimate the postclosure polymer
concentration, maximum allowable persulfate concentration and retained
permeability factor for the treatment design.

2.4.2 Summary of Software Features for the CleanFRAC Service Design


The FracCADE software features that involve the CleanFRAC Service Design
include the following features.
• The calculation of a retained fluid factor corresponding to an input breaker
concentration (for borate-crosslinked fluids only).
• A manual input of a retained fluid factor (for any fluid).
• The calculation of a postclosure polymer concentration within the fracture
proppant pack (for any fluid).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 8 of 20

• The FracNPV module to optimize breaker concentrations (for any fluid).


• Long-term proppant permeability values in the proppant editor.
• A time/temperature model to aid in scheduling breaker concentrations that
accounts for the cooldown effects of fracturing fluids.

2.4.3 Design Procedure Using a Manual Calculation


1. Divide the total proppant quantity (lbm) by the total fluid quantity (gal) to
determine the average proppant concentration for the treatment (lbm/gal).
2. Determine the polymer concentration factor for the treatment. Use Table 2,
Fig. 3 and the average proppant concentration (determined in Step 1).
3. Estimate the postclosure polymer concentration (lbm/1000 gal) by multiplying the
surface polymer concentration (lbm/1000 gal) by the polymer concentration
factor from step 2.
4. Determine the maximum allowable ammonium persulfate concentration for each
stage by using the rheology selection guide provided in Treatment Design and
fluid rheology data provided in the Fracturing Materials Manual  Fluids.
5. Determine the active breaker added by multiplying the total J475 added (lbm) by
0.8, and then adding the total J218 (lbm).
6. Determine the average active breaker concentration by dividing the pounds
active breaker added (Step 5) by the total fluid volume (Mgal) pumped (not
including flush).
7. Determine the ammonium persulfate breaker (lbm)/100 lbm of postclosure
polymer concentration by dividing the average active breaker concentration
(Step 6) by the postclosure polymer concentration (Step 3) and multiplying times
100.
8. Estimate the retained proppant-pack permeability by using the retained
permeability figures (Fig. 4 through Fig. 6) and the value determined in Step 7.

Table 2. Proppant-Pack Porosity of Sand and Intermediate-Strength Proppant


Porosity (%) Closure Stress Sand (psi) Closure Stress ISP (psi)
37.0 1000 4000
33.5 3000 6000
30.0 5000 8000

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 9 of 20

Fig. 3. Effects of proppant concentration and porosity on postclosure polymer


concentration.

Fig. 4. Influence of ammonium persulfate breaker on retained proppant-pack


permeability for linear fracturing fluids.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 10 of 20

Fig. 5. Influence of ammonium persulfate breaker on retained proppant-pack


permeability for borate-crosslinked fracturing fluids.

Fig. 6. Influence of ammonium persulfate breaker on retained proppant-pack


permeability for organometallic-crosslinked fracturing fluids.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 11 of 20

2.4.4 Design Procedure Using the FracCADE Software Exclusively or in Combination


With a Manual Calculation
1. Use the FracCADE software to determine the fracture design, the exposure time
that the fluid stage will be at BHST and a user input observation temperature.
2. Determine the maximum allowable ammonium persulfate breaker (J218)
concentration using the FracCADE exposure time at temperature prediction and
the fluid rheology data in the Fracturing Materials Manual  Fluids. The
fracturing fluid must maintain a minimum of 100 cp at 170 sec-1 during the time
the fluid is transporting proppant. In some cases, laboratory data with the job
mix-water is preferred to standard manual data. If field experience indicates that
higher J218 concentrations than those provided in the Fracturing Materials
Manual  Fluids can be used, then use the field experience as a guideline.
NOTE: The breaker rheology data provided in the Fracturing Materials Manual
 Fluids are considered to be conservative values because the concentration of
the polymer due to fluid leakoff is not considered.
3. Determine the expected percent premature release value for the J475 (Section
2.2.1 [Table 1]).
4. Determine the maximum allowable J475 concentration (lbm/1000 gal) for each
stage by dividing the maximum allowable J218 concentration (Step 2) by the
expected premature release value (Step 3, decimal equivalent).
5. Use the FracNPV analysis to determine the average J475 concentration that will
be the most cost effective. Do not exceed the maximum allowable amount
determined in Step 4.
6. Estimate the total quantity of J218 and J475 to be used. Use the average J218
concentration (Step 2), average J475 concentration (Step 5) and total fluid
volume. Schedule the J475 into the fluid stages so no stage receives less than
2 lbm/1000 gal unless the maximum allowable J475 concentration is less than
2 lbm/1000 gal. Maximize the J475 in the pad and early proppant stage.
7. Design a breaker schedule using the optimum average J475 concentration,
exposure time at temperature data and maximum allowable concentration of
J218. Determine the quantity of ammonium persulfate breaker (J218)
prematurely released from the J475. The ammonium persulfate breaker
released from the J475 must be less than or equal to the maximum allowable
determined in Step 2. Supplement the J475 with J218 so each fluid stage will be
exposed to the maximum allowable amount of ammonium persulfate breaker
(ammonium persulfate breaker prematurely released from J475 + additional
J218 = maximum allowable J218).
8. Determine the feeder metering rates. J475 is continuously mixed via a dry-
additive system, and J218 is mixed and metered as an aqueous solution.
provides instructions for the determination of feeder metering rates.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 12 of 20

3 Execution
Two metering devices have been developed to accurately meter breaker
concentrations during the job execution for the CleanFRAC Service. They may be
installed on the POD* blender. Liquid-additive metering systems are also installed on
the PCM* precision continuous mixer.
Auger-Type, Dry-Additive Feeder  An auger-type, dry-additive feeder is available
to continuously meter the J475 Breaker. Each mixer on the POD blender is
equipped with a dry-additive feeder.
Liquid-Additive Metering System  A liquid-additive metering kit for Breaker J218 is
available and is designed to pump J218 as an aqueous solution. All components of
the liquid-additive metering kit are manufactured from corrosion-resistant materials.
A detailed explanation of the metering devices installed on the POD blender is
provided in Section 12, Phase 1 Supplement of the SBT611 POD Blender Operator's
Manual.
A detailed explanation of the liquid-additive metering system installed on the PCM
precision continuous mixer is provided in the PCM Precision Continuous Mixer
Operator's Manual.

3.1 Metering the J475 Using the Auger-Type, Dry-Additive Feeder


The auger-type, dry-additive feeder is supplied with two augers to cover a wide
range of metering rates. The coarse pitch auger delivers approximately twice as
much per revolution as the fine pitch auger. The augers are designed to operate
within a range of 5 to 200 rpm. This corresponds to an additive rate of 0.7 to 50
lbm/min. The fine pitch auger should be chosen if either auger can be used for a
particular job. The volumetric displacement for the fine auger is 3.25 in.3/rev. The
volumetric displacement for the coarse auger is 5.5 in.3/rev. The bulk volume
specific gravity of the J475 is approximately 1.19. The empirically derived volumetric
displacements for J475 using the fine auger is 0.1398 lbm/rev. The empirically
derived volumetric displacement for J475 using the coarse auger is 0.2502 lbm/rev.
The minimum and maximum metering rates for both screws are provided in Table 3.
Table 3. J475 Metering Rates Auger-Type, Dry-Additive Feeder
Rate Fine Auger (lbm/min) Coarse Auger (lbm/min)
Minimum 0.7 1.25
Maximum 27.96 50.04
The procedure for scheduling the J475 using the auger-type, dry-additive feeder is
provided in. An example is provided in Section 4.3. This procedure (through step 4)
is automatically calculated by the FracCADE software.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 13 of 20

3.1.1 Procedure for Scheduling the J475 Using the Auger-Type Dry-Additive Feeder
1. Use a copy of the worksheet provided on page 14. Input the known information
into the header and table.
2. Determine the clean fracturing fluid pump rate for each stage (Eq. 1).
qd
qc =
( )
(1)
1 + C p / Dp

Where:
qc = clean fracturing fluid pump rate (bbl/min)
qd = fracturing fluid pump rate with proppant (bbl/min)
Cp = pounds proppant added per gal of fracturing fluid
Dp = proppant absolute density (lbm/gal).
3. Determine the J475 addition rate for each stage (Eq. 2).
q J 475 = 0.042 C J475 qc (2)
Where:
qJ475 = J475 addition rate (lbm/min)
CJ475 =J475 concentration (lbm/1000 gal)
qc = clean fracturing fluid pump rate (bbl/min).
4. Note the minimum and maximum calculated J475 addition rates. Use Table 3 to
determine the proper auger size to use. If either auger will work, use the fine
auger. Write down the auger selection and the corresponding volumetric
displacement. Volumetric displacement for the fine auger is 0.1398 lbm/rev; for
the coarse auger it is 0.2502 lbm/rev.
5. Determine the feeder rate for each stage (Eq. 3).
q J 475
q feed = (3)

Where:
qfeed = feeder rate (RPM)
qJ475 = J475 addition rate (lbm/min)
Sv = volumetric displacement (lbm/rev).

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 14 of 20

J475 metering worksheet (Using the auger-type dry-additive feeder)


Slurry Pump Rate (bbl/min) :
Proppant Size (mesh range) :
Proppant Density (lbm/gal) :
J475 Breaker Total (lbm) :
Minimum Additive Rate (lbm/min) :
Maximum Additive Rate (lbm/min) :
Auger Type (Fine, Course) :
Auger Rate (lbm/rev) :

Proppant Clean J475 Conc. J475 Total Additive Feeder


Stage Rate (lbm/1000 gal) (lbm) Rate Rate
(bbl/min) (lbm/min) (RPM)

3.2 Metering the J218 as an Aqueous Solution


The instructions for mixing and metering aqueous J218 solutions are provided in
Section 3.2.1. Additional information is provided in Treatment Execution. An
example is provided in Section 4.4. The metering values are automatically
calculated by the FracCADE software.

3.2.1 Procedure for Calculating Mixing and Metering Parameters for J218 Solutions
1. Use a copy of the worksheet provided on page 16. Input the known information
into the header and table.
2. Determine the maximum solubility (wt%) of J218 in water. Water solubilities of
J218 as a function of temperature are provided in.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 15 of 20

3. Based on the maximum solubility, determine the solution makeup (lbm J218/25
gal water). Use the mixing and metering parameters table for J218 provided in
Treatment Execution. Because of the low concentrations (lbm/1000 gal) of J218
that are often pumped, it may not be feasible to mix highly concentrated
solutions of J218. Since J218 is packaged in 11 lbm bags, round up to the
nearest multiple of 11. This will also provide a few additional gallons of solution
to compensate for possible inaccuracies in injection rate.
4. Determine the final solution volume.
5. Determine the solution metering rate (gal/1000 gal) for each stage. Divide the
J218 concentration (lbm/1000 gal) in the fracturing fluid by the weight of J218
per gallon of solution (from the J218 mixing and metering parameters table).
6. Determine the solution injection rate for each stage. Use Eq. 4 when the fluid
does not contain proppant. Use Eq. 5 when the fluid contains proppant.
qiJ 218 = 0.042qmJ 218 qc (4)
Where:
qiJ218 = J218 solution injection rate (gal/min)
qmJ218 = J218 solution metering rate (gal/1000 gal of fracturing fluid)
qc = clean fracturing fluid pump rate (bbl/min).
0.3503 qmJ218 qd SG p
qiJ 218 = (5)
8.32 SG p + C p

Where:
qiJ218 = J218 solution injection rate (gal/min)
qmJ218 = J218 solution metering rate (gal/1000 gal of fracturing fluid)
qd = fracturing fluid pump rate with proppant (bbl/min)
SGp = proppant specific gravity
Cp = pounds proppant added per gal of fracturing fluid.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 16 of 20

J218 metering worksheet (Using the liquid-additive system)


Slurry Pump Rate (bbl/min) :
Proppant Size (mesh range) :
Proppant Specific Gravity (SGp) :
J218 Breaker Required (lbm) :
Surface Fluid Temperature (°F[°C]) :

Proppant Pump J218 Solution Solution Stage Stage


Stage Rate Concentration Metering Rate Injection Time Solution
(bbl/min) (lbm/1000 gal) (gal/1000gal) Rate (min) Total
(gal/min) (gal)

4 Examples

4.1 Example Design Using a Manual Calculation


Design Parameters 
YF140 = 418,000 gal
Proppant = 1,110,000 lbm 20/40-mesh sand (33.5% porosity)
Breaker J475 = 3900 lbm
Breaker J218 = 250 lbm
1. Determine the average proppant concentration.
1,110,000 lbm
= 2.7 lbm / gal
418,000 gal
2. Determine the polymer concentration factor  Refer to Fig. 3.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 17 of 20

Assume porosity = 33.5%. For a proppant concentration of 2.7 lbm/gal, the


polymer concentration factor is approximately 16.
3. Estimate the postclosure polymer concentration.
(40 lbm/1000 gal)(16) = 640 lbm/1000 gal
4. Determine the Maximum Allowable J475 Concentration for Each Stage.
Maximum Allowable J218 Concentration
Maximum = (lbm / 1000 gal )
J 475 Expected Premature Release Value
( decimal )
5. Determine the active breaker added.
(3900 lbm)(0.8) + 250 lbm = 3370 lbm
6. Determine the average active breaker concentration.
3370 lbm
= 8.1 lbm / 1000 gal
418 Mgal
7. Determine the concentration of ammonium persulfate breaker (lbm)/100 lbm of
postclosure polymer concentration.

8.1 lbm / 1000 gal


(100) = 1.27
640 lbm / 1000 gal
8. Estimate the retained proppant-pack permeability (refer to Fig. 5).
For 1.27 lbm of ammonium persulfate breaker/100 lbm of postclosure polymer
concentration and a temperature of 152°F (67°C), the retained proppant-pack
permeability is estimated to be 21% of the clean proppant-pack permeability.

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 18 of 20

4.2 Example Procedure for Scheduling the J475 Using the Auger-Type Dry-Additive
Feeder
Slurry Pump Rate (bbl/min) : 40
Proppant Size (mesh range) : 20/40
Proppant Density (lbm/gal) : 22.1
J475 Breaker Total (lbm) : 3892
Minimum Additive Rate (lbm/min) : 12.4
Maximum Additive Rate (lbm/min) : 15.4
Auger Type (Fine, Course) : Fine
Auger Rate (lbm/rev) : 0.1398

Proppant Clean J475 Conc. J475 Total Additive Feeder Rate


Stage Rate (lbm) Rate (RPM)
(lbm/1000 gal)
(bbl/min) (lbm/min)
PAD 40.0 8 1112 13.4 95.9
2 PPA 36.7 10 1100 15.4 110.2
4 PPA 33.9 10 830 14.2 101.6
6 PPA 31.5 10 610 13.2 94.4
8 PPA 29.4 10 240 12.4 88.7

1. Determine the clean pump rate for each stage. For the 4 PPA stage 
40
qc = = 33.9 bbl / min
1+ ( )
4
22.1

2. Determine the J475 addition rate for each stage. For the 4 PPA stage 
q J 475 = 0.042 (10) (33.9) = 14.2 lbm / min
3. Determine the proper auger size. The minimum J475 addition rate is
12.4 lbm/min and the maximum rate is 15.4 lbm/min. These rates are within the
range of either auger. If either auger will work, the fine auger should be selected.
Volumetric displacement for the fine auger is 0.1398 lbm/rev.
4. Determine the feeder rate for each stage. F or the 4 PPA stage 
14.2
q feed = = 101.6 RPM
0.1398

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFRAC Service
Page 19 of 20

4.3 Example Procedure for Calculating Mixing and Metering Parameters for J218
Solutions
Slurry Pump Rate (bbl/min) : 40
Proppant Size (mesh range) : 20/40
Proppant Specific Gravity (SGp) : 2.65
J218 Breaker Required (lbm) : 125
Surface Fluid Temperature (°F[°C]) : 60(16)

Proppant Pump J218 Solution Solution Stage Stage


Stage Rate Concentration Metering Rate Injection Rate Time Solution
(bbl/min) (gal/1000 gal) (gal/min) (min) Total (gal)
(lbm/1000 gal)
PAD 40    83 
2 PPA 40    72 
4 PPA 40 0.2 0.465 0.66 59 38.9
6 PPA 40 1.0 2.33 3.08 46 141.7
8 PPA 40 2.0 4.65 5.74 19 109

1. Determine the maximum solubility (wt%) of J218 in water from the J218 water
solubility table provided in Treatment Execution. The maximum solubility (wt%)
of J218 at 60°F (16°C) is 39.1%.
2. Determine the solution makeup (lbm J218/25 gal water) using the J218 mixing
and metering table provided in Treatment Execution. Any J218 solution up to
132 lbm J218/25 gal water can be used. However, low J218 concentration in the
4 PPA stage would require an impractical solution injection rate if the solution
were mixed at 132 lbm J218/25 gal water. A more realistic solution makeup is
11 lbm J218/25 gal water. The total weight of J218 required is 125 lbm. Since
J218 is packaged in 11 lbm bags, it will be convenient to use 12 bags or 132
lbm. This will also provide a few additional gallons of solution to compensate for
possible injection rate inaccuracies.
3. Determine the final solution volume. The final solution volume is 306 gal (each
25 gal of water containing 11 lbm J218 produces a solution volume of 25.5 gal
and 12 volumes of 25.5 gal each equals 306 gal).
4. Determine the solution metering rate (gal/1000 gal) for each stage. For the
4 PPA stage 
0.2
qm 218 = = 0.465 gal / 1000 gal
0.43

DOWELL CONFIDENTIAL
Section 700.4
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFRAC Service Dowell
Page 20 of 20

5. Determine the solution injection rate (gal/min) for each stage. For the 4 PPA
stage 

qi218 =
( 0.3503)( 0.465)( 40)( 2.65) = 0.66 gal / min
8.32 ( 2.65) + 4

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 1 of 22

FRACTURE-HEIGHT-CONTAINMENT SERVICES

1 Introductory Summary............................................................................................................. 2
1.1 Fracture Height Prediction ................................................................................................... 3
1.2 Fracture Penetration ............................................................................................................ 3
1.3 Fracture Evolution................................................................................................................ 4
1.4 Fracture Height-Growth and Containment ........................................................................... 7
1.5 Fracture-Height-Containment Services................................................................................ 9

2 DIVERTAFRAC Service.......................................................................................................... 10
2.1 Discussion.......................................................................................................................... 10
2.2 Design Methodology .......................................................................................................... 11
2.2.1 Design Example........................................................................................................ 13
2.3 Execution Methodology...................................................................................................... 15
2.4 DIVERTAFRAC Fluid Using Water Control Agent S41 as a Diverting Material ................. 15
2.4.1 Spacers .................................................................................................................... 15
2.4.2 Dilution with Sand ..................................................................................................... 15
2.4.3 Additional Information ............................................................................................... 15

3 INVERTAFRAC Service.......................................................................................................... 16
3.1 Design Methodology .......................................................................................................... 16
3.2 Execution Methodology...................................................................................................... 17
3.2.1 Field Mixing Procedures ........................................................................................... 18
3.2.2 Additional Mixing Techniques ................................................................................... 19

4 Computer-Aided Design........................................................................................................ 20
4.1 Additional Computer-Aided Job Design Information .......................................................... 22
FIGURES
Fig. 1. First phase of evolution..................................................................................................... 4
Fig. 2. Illustration of fracture growth............................................................................................. 5
Fig. 3. Linear plot of pressure ...................................................................................................... 6
Fig. 4. Log-Log plot of net pressure. ............................................................................................ 6
Fig. 5. Pressure and width for growth through barriers................................................................ 7

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 2 of 22

Fig. 6. Height control and conventional treatment for an offset well.............................................8


Fig. 7. Principles of fracture-height-containment techniques........................................................9
Fig. 8. J423 rise rate versus fluid viscosity (static conditions). ...................................................17
Fig. 9. J423 flow rate versus area. .............................................................................................19
TABLES
Table 1. J423 Addition Rate .......................................................................................................18

1 Introductory Summary
The fracturing process is governed by the net pressure. The net pressure is the
pressure above the closure pressure, that is, piw-pc, where piw is the reservoir
fracturing pressure and pc is the closure pressure. The equations for interpreting the
pressure result from combining the basic relations of material balance, fluid flow, and
fracture compliance. Eq. 1 defines the net pressure relationship for the PKN model.
Eq. 2 defines the net pressure relationship for the KGD model. Eq. 3 defines the net
pressure relationship for the Radial model.
1/ 2n' +2
 
(
∆p f ∝ E 2n′+1K ′qi )
n′ 1/ 2n' +2

xf

 h f 3n' +1 
(1)

1/ 2n' +2
1/ 2n' +2  
(
∆p f ∝ E 2n′+1K ′qi n′ ) 
n′
1

 h f x f 2n' 
( 2)

1/ 2n' +2  1  1/ 2n' +2
(
∆p f ∝ E 2n′+1K ′qi
n′
)  R 3 n' 
 
( 3)

Where:
∆pf = net pressure (psi)
E = modulus of elasticity
n′ = behavior index
K′ = consistency index
qi = injection rate (bbl/min)
xf = fracture half-length (ft)
hf = fracture height (ft)
R = fracture extension (ft).
These relationships for pressure indicate that their dependence on fluid rheology,
injection rate and modulus are the same for all models, and that the dependence on
fracture extension and height is different. For increasing penetration, the net

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 3 of 22

pressure increases for the PKN model and decreases for the KGD and Radial
Models.
As the net pressure approaches the stress difference of a barrier, the fracture
penetrates the barrier. The in-situ stress level of the barrier will limit fracture height
but not confine the fracture and, therefore, fracture width development in the barrier
is restricted. The restriction will act as a bridging point for an artificial low-
permeability barrier created with a diverting material.

1.1 Fracture Height Prediction


Fracture height prediction methods were at best qualitative until the late 1970s. Until
then, a common method but a misconception was to use the spontaneous potential
(SP) and gamma ray (GR) curves to identify shale from sand. The location of an
arbitrarily thick shale was considered to be the limit of the vertical extent, because
shale formations generally have higher stress than reservoir formations and the
interface between the shale and sand would slip and cause blunting. Another
common misconception, (still used today), is that shale is an absolute barrier to
fracture migration.
Several investigators have used stress profiles and rock properties in linear fracture
mechanics models to predict fracture height. Stress data were obtained from mini-
or microfrac tests. Although this is the preferred approach when the data are
reliable, the frequency of the data, especially for shales, is scanty.
Rock properties can be inferred from log-derived compressional and shear velocities.
The introduction of long-spaced sonic and sonic digital tools during the late 1970s
and early 1980s improved the reliability of these measurements, and allowed the
calculation of stress and mechanical properties for every six inches over the entire
logging depth. This improved fracture height predictions; however, the inferred
stresses are approximate and primarily provide qualitative predictions.
Fracture height prediction includes the following.
• Compressional and shear wave slowness from a long-spaced sonic or sonic
digital tool to calculate rock elastic properties.
• A transversely elastic model to compute minimum horizontal stress for each
layer.
• Calibration of the log derived stresses with actual measured stresses in selected
formations.
• A linear fracture mechanics model to predict fracture height for corresponding
pressures.

1.2 Fracture Penetration


Hydrocarbon production increases are directly related to fracture penetration of the
producing interval, unless production is fracture-conductivity-limited. Fracture

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 4 of 22

containment in the producing zone increases fracture penetration and provides


greater production increases. If a good fracture conductivity ratio can be achieved,
then a fracture penetrating 100% of the drainage radius can provide more than a
13-fold increase in production.

1.3 Fracture Evolution


The following figures illustrate the fracture geometry for an ideal case with bounding
formations of higher stress. In Fig. 1, the fracture is in the first phase of evolution.
Bounding formations with higher stress provide the primary barrier to fracture height-
growth and permit efficient fracture extension. A fracture in its initial phase of
propagation from a perforated interval will increase as expanding ellipses if the
fracture is initiated from a line source (perforated interval). A fracture in its initial
phase of propagation from a point source (perforated section less than the bounded
height or a fracture inclined from the wellbore) will increase as a series of expanding
circles. The duration of the first stage may be a very short time (minutes) in a
relatively small zone. In a massive zone the first stage may exist for the entire
treatment.

Fig. 1. First phase of evolution.


Fig. 2 illustrates the shape of fracture growth in a propagating vertical fracture with
barriers above and below. In this illustration, the barriers have a higher fracturing
pressure (greater stress contrast) than the zone of interest.
Three phases of fracture growth are shown. Phase 1 indicates radial circular
fracture growth until the barriers are reached. In Phase 2, the barriers force the
fracture from its preferred circular shape into a radial lateral shape between the
barriers. The pressure required to propagate the fracture between the barriers will
progressively increase. As this increasing pressure reaches the fracturing pressure
of the barriers, the barriers will rupture and the fracture height will increase into the
barrier formations. This is shown as Phase 3. The fracture height is dependent to a
large degree on the hydraulic fracturing pressure of the zones, both above and

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 5 of 22

below the zone of interest. In both barriers have equal fracturing pressure. In
practice, this may or may not be the case and fracture height can extend above or
below (or both) the zone of interest.

Fig. 2. Illustration of fracture growth.


Fig. 3 and Fig. 4 represent plots of the relative bottomhole pressure during each of
the three phases shown in Fig. 2. The numbers indicate either Phase 1, Phase 2, or
Phase 3. Fig. 3 is a linear plot of pressure versus time (or volume). In Fig. 4,
however, a log-log plot of pressure versus time becomes much more meaningful.
Phase 1 represents circular radial fracture growth and is shown in Fig. 4 as negative
slope. Phase 2, representing radial lateral fracture extension between barriers,
indicates a positive log-log slope of 1/8 or 1/4 depending upon the fluid and its
efficiency. Phase 3 is indicative of little or no fracture growth, due possibly to leakoff
(pump rate and leakoff being approximately equal at this phase). In this no-growth
phase, the fracture is filling with fracturing fluid and proppant and screenout may
occur.
In Phase 3, the pressure in the zone of interest between the barriers is slowly
increasing. If the log-log slope of Phase 3 changes, it can be due to fracturing out of
zone (unrestricted fracture height-growth Phase 1) or to further propagation of the
fracture between barriers (Phase 2).

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 6 of 22

Fig. 3. Linear plot of pressure.

Fig. 4. Log-Log plot of net pressure.


In summary, decreasing net pressure during the initial growth period can be
interpreted as an indication of the unrestricted growth of a radially evolving fracture
in a vertical or horizontal plane. This unrestricted growth is defined as Phase 1.
Phase 2 is shown as increasing net pressure with a small log-log slope after the
initial growth period is indicative of a vertical fracture extending primarily in length
with restricted height growth. Phase 3, with pressure constant, indicates no fracture
growth.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 7 of 22

1.4 Fracture Height-Growth and Containment


The pressure and vertical cross section of the width profile for growth through a
stress barrier are illustrated in Fig. 5. Stage a and Stage b are the same as for
growth into a barrier shown by Phase 2 and Phase 3 in Fig. 4. The initial radial
Phase 1 is not shown. Stage c results when the barrier is crossed and the fracture
enters a lower stress zone. When this occurs, the fluid pressure is greater than the
stress of the zone and results in an accelerated rate of growth at decreasing
pressure and width in the primary zone, that is, unstable height growth. The
unstable growth begins at the well where the pressure is greatest and will progress
further along the fracture as pumping continues. In a similar manner, the unstable
growth will cease progressively along the fracture if and when another stress barrier
is encountered. If another barrier is encountered, the pressure will begin to increase
again.

Fig. 5. Pressure and width for growth through barriers.


A decreasing pressure profile generally indicates unconfined height-growth in the
case of a radial fracture. The width profile indicates that a pinch-point occurs in the
barrier during Stage c. The pinch-point can cause proppant to bridge at this location
as fluid is permitted to pass through. Consequently, excessive dehydration of the
slurry and decreasing width in the primary zone can result in a rapid screenout even
with sufficient pad volume and at low proppant concentrations. The slurry
dehydration, decreasing width, and height growth can be significantly reduced by
placing a relatively impermeable mixture of diverters in the fracture prior to the main
fracturing treatment.
Fracturing treatments with foam fracturing fluids can also retard height growth
because of the yield stress for these fluids (see Foam Fracturing). The yield stress
permits a pressure gradient to be developed which is proportional to the yield stress
and inversely proportional to the width which is very small at the pinch-point.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 8 of 22

The upper portion in Fig. 6 shows the log-log plot of net pressure for a treatment with
a significant period of growth through a barrier. Growth through a barrier is indicated
by the decreasing pressure following a period of increasing pressure as illustrated in
Fig. 5. The upper portion of Fig. 6 indicates a rapid screenout near the wellbore
(log-log slope greater than unity) after the addition of proppant. The lower portion of
shows a subsequent and successful treatment of an offset well using an intermediate
stage of diverters (denoted as “nonproppant” on the figure). This treatment also
used less fluid volume and fluid viscosity to reduce the net pressure and rate of
height growth. The pressure response for the treatment shows no indication of
growth through a barrier (decreasing pressure) and shows an increase in the
pressure after the diverter was injected and bridged in the pinch-point. This
pressure increase resulted because the fracture began responding as if the height
growth had not occurred, that is, extrapolated response from Stage a of Fig. 5.

Fig. 6. Height control and conventional treatment for an offset well.


Fracture height growth is generally governed by higher stresses in the boundary
formations. Growth into the barrier depends on the ratio of net pressure within the
fracture and the stress difference between the reservoir and barrier. For a ratio of
about 0.4, the height growth is negligible. For a ratio of about 0.65, the total fracture
height is twice the reservoir height and would require each barrier thickness to be at
least one-half the height of the reservoir. For a ratio of about 0.8, the height is three
times that of the reservoir and would require each barrier thickness to be at least
equal to the height of the reservoir. Therefore, the barrier requirements depend on
the thickness of the reservoir and the magnitude of the net pressure. Net pressure
has an approximate inverse proportional relation with height. The net pressure will
approximately double when the reservoir height is halved. The general conclusion is
that a smaller zone is more likely to experience height growth and will require both
higher stress differences and thicker barrier formations for height confinement.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 9 of 22

1.5 Fracture-Height-Containment Services


The Dowell fracture-height-containment services are
• the DIVERTAFRAC* Service
• the INVERTAFRAC* Service
• the BRACKETFRAC* Service.
Simultaneous application of the DIVERTAFRAC and INVERTAFRAC services is the
BRACKETFRAC Service.
Applications
Fracture-height-containment services can be used in all fracturing operations where
fracture height-growth out of the productive zone is predicted or experienced.
• Weak Zones  If the in-situ stress levels of the barrier zones are less than, equal
to or slightly greater than the pressure required to propagate the fracture, the
fracture will grow in height through the barrier zones.
• Undesirable Zones  In addition to barren zones, water zones or a gas cap (oil
wells) may exist. Fracturing into a water zone may result in increased water
production which could limit production and increase disposal costs. Fracturing
into a gas zone could stimulate gas production instead of oil production, resulting
in premature depletion of reservoir energy and reduced ultimate recovery.
The principles of these fracture-height-containment techniques are illustrated in
Fig. 7.

Fig. 7. Principles of fracture-height-containment techniques.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 10 of 22

2 DIVERTAFRAC Service
Undesired fracture migration downward can be controlled with the DIVERTAFRAC
Service. This service places a solid diverting material (usually sand) in the bottom of
a fracture. The diverting material creates an artificial barrier of reduced permeability
resulting in fracture growth upward and outward.
Where a water-bearing zone exists below a hydrocarbon zone, a polymer-coated
sand (Water Control Agent S41) may be used as the diverting material. This
material is transported in an oil-base fluid. The polymer coating swells and fills the
pore space between the sand grains when it is contacted by encroaching water or
brine.
Gas Wells with Water Contact
Some gas reservoirs have bottom water support through rechargeable aquifers
leading to volumetric displacement. Often in such cases, the water influx may be
accelerated with the improvement in vertical permeability caused by hydraulic
fracturing. The DIVERTAFRAC Service using S41 can be effective in controlling
water production.

2.1 Discussion
Fracturing oil reservoirs containing bottom water often resulted in excessive water
production. Patented in the 1960s, the DIVERTAFRAC Service was offered as a
method to prevent undesired water production.
The initial technique used in the DIVERTAFRAC Service placed polymer-coated
diverting material (polymer-coated sand) in the bottom of a fracture. A low-viscosity,
oil-base fluid was used to create the fracture and transport the diverting material.
The diverting material settled at the bottom of the fracture and created an artificial
barrier which would prevent downward fracture growth. The polymer, designed to
swell when contacted by encroaching water, would expand and fill up the pore space
if the fracture penetrated the water zone. The reduced permeability caused by the
swelling of the polymer prevented water migration through the fracture. An absence
of increased water production after the treatment indicated a successful application.
The possibility of plugging highly water-saturated, hydrocarbon-producing zones and
the requirement for water-free oil as a carrier fluid were disadvantages to this
application.
The same technique was used in formations that were tens of feet above the water
table using uncoated (plain) sand as a substitute for polymer-coated sand. This
diversion technique relied on proppant banking prior to reaching the water table and
proved to be successful in preventing excessive water production after treatments.
The DIVERTAFRAC procedure became more complex with the advent of massive
hydraulic fracturing because it required placement of the diverting material further
away from the wellbore.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 11 of 22

2.2 Design Methodology


1. Design the pad for the DIVERTAFRAC Service. The fluid and pump rate should
be the same as the main fracturing treatment. The pad volume should be
sufficient to produce a radius equal to the distance from the mid or bottom
perforations to the weak or undesirable zone.
2. Determine the required fluid viscosity. The diverter should settle a distance
equivalent to the fracture radius (determined above) in 10 min. Settling velocity
(Stoke's law) for a Power-law fluid may be determined using Eq. 4. Settling
velocity for a Newtonian fluid may be determined using Eq. 5.
1/ n ′
60( 2n ′ + 1) D  (ρ p − ρ f )7.48 D 
Vs =   ( 4)
108 n'  72K ′ 
8246 D 2
Vs = (ρ p − ρ f ) ( 5)
µ
Where:
Vs = terminal settling velocity (ft/min)
n′ = behavior index
K′ = consistency index
ρp = diverter absolute density (lbm/gal)
ρf = fluid absolute density (lbm/gal)
D = average diverter diameter
µ = fluid viscosity (cp).
3. To stop downward propagation, determine the diverter pack-height and
permeability necessary to achieve a 100 psi (minimum) pressure differential. A
10 ft pack-height and permeability approximately 1000 md is a good starting
point. The diverter particle size should be as wide a range as possible for the
lowest permeability. However, if the diverter mesh-range is too small, the diverter
fall-rate may be too slow. Diverter-pack permeability can be determined using
Eq. 6. The pressure differential can be determined using Eq. 7.
2
  D 
k = ( 25, 400)   (6)
  x 

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 12 of 22

Where:
k = diverter-pack permeability (md)
D = average particle size (in.)
x = a factor which depends on mesh range
x = 2.36 for 40/60 mesh
x = 2.00 for 20/40 mesh
x = 2.30 for 16/20 mesh
x = 3.3 for 100 mesh (assumed)
x = 6 for 100 mesh and 40/60 (assumed).
D
Note: can be replaced by the average effective pore size (in.).
x
 C µh 2 
∆p = 273E 6  w d  (7)
 tkw 
Where:
∆p = pressure differential in the diverter pack (psi)
Cw = formation fluid-loss coefficient (ft/min1/2)
µ = fluid viscosity (cp)
hd = diverter-pack height (ft)
t =time (min)
k = diverter-pack permeability (md)
w =fracture width at the top of the diverter pack (in.).
4. Determine the volume of diverter required. Use the designed pack-height and a
pack-length equal to 1½ times the radius (determined in Step 1).
5. Determine the carrier fluid volume. Use a concentration of one lbm of diverter
material/gal fluid. If leakoff will be severe, use more fluid. If leakoff will be low,
use less fluid.
6. Design the main fracturing treatment. Design for a horizontal penetration of two
times the radius (determined in Step 1). Plan to have the pad depleted at the end
of the treatment so that unnecessary penetration is avoided. This risks a tip
screenout, which is acceptable if the pressure is not allowed to increase too high.
Determine a fracture height assuming the bottom of the fracture is one radius
(determined in ) from mid-perforation. Use a constant-height fracture model. If
upward fracture growth is possible, use a semi-radial model (radial model with
twice the rate and volume.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 13 of 22

Design Notes
This design method was developed for, and used in, North Sea operations. A
classical case in which a DIVERTAFRAC treatment was performed (with water 60 ft
from the bottom perforations) had no water production. Diverting-type treatments
performed by a competitor in an offset well resulted in mainly water production.
This method varies slightly from the FracCADE* method which concentrates on how
much net pressure can be tolerated along the fracture. The implicit assumption in
this method is that the pack needs to be placed from the wellbore to at least 50%
penetration and preferably 75% penetration.

2.2.1 Design Example


• Reservoir Parameters
• perforated interval = 55 ft
• water table is 27 ft below the bottom perforation
• YF*640LT fluid is selected as the main fracturing fluid
• fracture length less than 100 ft.
1. The distance from the bottom perforation to the water is 27 ft. Using this number
(27 ft) as the fracture radius is safe, but is very small and the job will be difficult
to pump.
The distance from mid-perforation to the water is 55 ft. This number (55 ft) can
be used for the fracture radius if conditions are ideal and the maximum fracture
length is desired.
In this case, a compromise of 40 ft for the fracture radius is used. With diverting
material placed 40 to 60 ft, a final fracture length of 80 ft can be designed. The
FGS simulator in the FracCADE software indicates that 20 bbl of YF640LT will
be required to create an initial fracture (wall-building fluid-loss coefficient
= 0.0023 ft/min½, pump rate = 5 bbl/min).
2. Determine the required fluid viscosity. For a fracture radius of 40 ft, the diverting
material should settle at 4 ft/min. Estimating a fluid viscosity of 5 cp and using
Eq. 5, a settling velocity of 4.1 ft/min is calculated.
8246( 0.0135) 2
Vs = ( 22.1 − 8.34)
5
Vs = 4.1 ft / min

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 14 of 22

3. Determine the diverter permeability and pack-height. Use the following


parameters.
Cw = 0.0023 ft/min1/2
µ = 5 cp
hd = 5 ft
t = 20 min
w = 0.05 in.
D = 0.0135 in.
Using Eq. 6
2
  0.0135  
k =  25,400 
  2.36  
k = 21,111 md
Using Eq. 7
 ( 0.0023)( 5)( 5) 2 
∆p = 2,730,000  
 20 ( 21,111)(0.05) 
∆p = 166 psi
4. Determine the diverter volume (fracture length, 120 ft [two, 60 ft wings], pack
height, 5 ft, width, 0.05 in.).
(120)( 5)(0.05)
Volume =
12
Volume = 2.5 ft 3
At a bulk density of 100 lbm/ft3, 250 lbm of diverter will be required.
5. Determine the carrier fluid volume. At one lbm/gal, this requires six bbl of carrier
fluid (fluid viscosity, 5 cp).
6. Design the main fracturing treatment for a 80-ft fracture half-length.
The fluid volumes in this case are small. The fracturing fluids may need to be
spotted since bullheading a tubing volume might create a fracture that contacts
the water.
Example Notes
An assumption is made that once the DIVERTAFRAC pad and treatment are placed,
there is a pack (5 or 10 ft) of diverting material at the bottom tip of the fracture. For
this pack to stop downward propagation, a pressure drop of at least 100 psi
(vertically) is required. This pressure drop depends on vertical flow through the
pack. Since the bottom tip is not propagating, all the flow is due to fluid loss from the
pack to the formation. This is why the formation fluid-loss coefficient is used in the
design.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 15 of 22

2.3 Execution Methodology


1. Pump the DIVERTAFRAC pad fluid at the same rate as the main fracturing
treatment.
2. Pump the first 2/3 of the carrier fluid volume at the same rate as the main
fracturing treatment.
3. Pump the remaining 1/3 of the carrier fluid volume at a pump rate equal to the
leakoff rate to keep the fracture open while the majority of the diverter is settling.
Leave some diverter in suspension to be pushed further along the fracture once
the main fracturing treatment pad has started.
4. Pump the first “x” barrels of the main fracturing fluid pad so that “x” plus the fluid
volume in step 3 give a pumping time of 10 min. Pump the fluid at the same rate
as the main fracturing treatment.
5. Continue pumping the main fracturing treatment at the designed injection rate.

2.4 DIVERTAFRAC Fluid Using Water Control Agent S41 as a Diverting Material
S41 is sand that has been coated with a synthetic polymer. The S41 is mixed in a
fracturing fluid composed of water-free oil. S41 falling rate and hence fluid viscosity
are critical for correct DIVERTAFRAC placement.
Temperature or pressure have no effect on the swelling properties of S41. The
artificial barrier is stable to 210°F (99°C) and 1500-psi differential pressure
indefinitely.

2.4.1 Spacers
If the DIVERTAFRAC Service is to be preceded or followed by a water-base fluid or
acid, an oil spacer should be used to prevent premature swelling of the S41.

2.4.2 Dilution with Sand


The polymer coating on S41 is designed to swell and plug the pore spaces. Dilution
with uncoated fracturing sand will reduce the effectiveness of S41.

2.4.3 Additional Information


The Fracturing Materials Manual  Additives provides additional information on
DIVERTAFRAC fluid containing S41.

3 INVERTAFRAC Service
Undesired fracture migration upward can be controlled with the INVERTAFRAC
Service. This service places a buoyant diverting material (Diverter J423) in the top
of a fracture. J423 may be mixed and pumped in any water-base, oil-base, or acid-

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 16 of 22

base fluid. J423 creates an artificial barrier of reduced permeability resulting in


fracture growth downward and outward.

3.1 Design Methodology


1. Design the pad for the INVERTAFRAC Service. The fluid and pump rate should
be the same as the main fracturing treatment. The pad volume should be
sufficient to produce a radius equal to the distance from the mid or top
perforations to the weak or undesirable zone.
2. To stop upward propagation, determine the J423 pack-height necessary to
achieve a 100 psi (minimum) pressure differential. J423 pack-heights are
typically 40 to 100 ft. The pressure differential can be determined using Eq. 8.

Cw µhd
2
∆p = (8)
3600 tw
Where:
∆p = pressure differential in the diverter pack (psi)
Cw = wall-building coefficient (ft/min1/2)
µ = fluid viscosity (cp)
hd = diverter-pack height (ft)
t = time (min)
w = fracture width at the top of the diverter pack (in.).
3. Determine the required carrier fluid viscosity. The J423 should rise a distance
equivalent to the fracture radius (determined in Step 1) as quickly as possible.
The rate that the J423 rises is dramatically affected by the viscosity of the carrier
fluid. The fluid viscosity will generally be less than 10 cp at 170 sec-1. Fig. 8
shows the rise rate of J423 versus the viscosity of the fluid.
4. Determine the volume of J423 required. Use the designed pack-height and a
pack-length equal to 1 1/2 times the radius (determined in Step 1).
5. Determine the carrier fluid volume. The J423 concentration may be 0.25 to
2.0 lbm/gal. Typically the concentration will range from 0.75 to 1.0 lbm/gal and
is pumped at a single concentration. If leakoff will be severe, use more fluid. If
leakoff will be low, use less fluid.
6. Design the main fracturing treatment. Design for a horizontal penetration of two
times the radius (determined in Step 1). Plan to have the pad depleted at the
end of the treatment so that unnecessary penetration is avoided. This risks a tip
screenout, which is acceptable if the pressure is not allowed to increase too
high. Determine a fracture height assuming the bottom of the fracture is one
radius (determined in Step 1) from mid-perforation. Use a constant-height
fracture model. If upward fracture growth is possible, use a semi-radial model

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 17 of 22

(radial model with twice the rate and volume.

Fig. 8. J423 rise rate versus fluid viscosity (static conditions).

3.2 Execution Methodology


1. Pump the INVERTAFRAC Service pad at the same rate as the main fracturing
treatment.
2. Pump the first 2/3 of the carrier fluid volume at the same rate as the main
fracturing treatment.
3. Pump the remaining 1/3 of the carrier fluid volume at a pump rate equal to the
leakoff rate to keep the fracture open while the majority of the J423 is rising.
Leave some J423 in suspension to be pushed further along the fracture once the
main fracturing treatment pad has started.
4. Continue pumping the main fracturing treatment at the designed injection rate.

3.2.1 Field Mixing Procedures


J423 is mixed in the same manner as propping agents; however, standard Dowell
density measuring systems cannot be used to monitor the J423 concentration. The
J423 concentration must be controlled volumetrically. Because of the low specific
gravity of J423 (0.6 to 0.7), the J423 concentration will be considerably more
material volumetrically than an equivalent concentration of propping agent. The
desired J423 addition rate corresponds to a certain size area and is controlled by

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 18 of 22

setting the gate opening on the blender or hopper. The procedure for determining
the J423 addition rate and size of area needed is as follows.
1. Determine the J423 addition rate from Table 1.
Table 1. J423 Addition Rate
Pump J423 Concentration (pounds added)
Rate,
bbl/min
0.25 0.50 0.75 1.00 1.25 1.50 2.00
1 10 20 28 35 43 50 62
2 20 40 56 71 86 100 124
3 30 60 83 106 129 150 185
4 40 80 111 141 172 200 247
5 50 100 139 177 215 250 309
6 60 120 167 212 257 300 371
7 70 140 195 247 300 350 433
8 80 160 222 282 343 400 494
9 90 180 250 318 386 450 556
10 100 200 278 353 429 500 618
11 110 220 306 388 472 550 680
12 120 240 334 424 515 600 742
13 130 260 361 459 558 650 803
14 140 280 389 494 601 700 865
15 150 300 417 530 644 750 927
16 160 320 449 565 686 80 989
17 170 340 473 600 729 850 1050
18 180 360 500 635 772 900 1112
19 190 380 528 671 815 950 1174
20 200 400 556 706 858 1000 1236
Example: At a fluid pump rate of five bbl/min, a J423 concentration of 0.75 lbm
added to a gallon of fluid would be a J423 addition rate of 139 lbm/min.

2. Using the J423 addition rate, the area required (in.2) to obtain the J423 addition
rate is selected from Fig. 9.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 19 of 22

Fig. 9. J423 flow rate versus area.


Example: At a J423 addition rate of 139 lbm/min, an area of six square inches would
be required.

3.2.2 Additional Mixing Techniques


• Insufficient fluid turbulence will allow the J423 to float on the surface of the fluid.
When using a paddle-type mixer to blend and mix J423, the paddles must be
turning at a speed that will ensure that fluid turbulence is maintained.
• Fluid must not be splashed into the gate opening. Wet J423 will reduce or stop
the flow of J423 through the gate.
• The fluid pump rate must be sufficient to suspend the J423 in the fluid as it
travels through discharge hoses and pumps. Use as few hoses and pumps as
possible to maintain high fluid velocity. Do not exceed maximum pump rates for
hoses and pumps.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 20 of 22

4 Computer-Aided Design
The BRACKETFRAC module is in the FracCADE software. The module allows a
variety of well parameters and treatment combinations to be simulated until the
optimum design is obtained. Computer simulation is available for the
DIVERTAFRAC Service, INVERTAFRAC Service, or BRACKETFRAC Service. Data
for the module can be obtained from well tests, repeat formation tests, sonic or
digital sonic logs, cores, and the DataFRAC* Service. Properties of fracturing fluids
are provided in the Fracturing Materials Manual  Fluids.
Benefits of the BRACKETFRAC Software are:
1. Minimize the amount of input data required by the engineer to obtain an estimate
of a job design for one of the fracture-height-containment services.
2. Generate a reasonable estimate of the requirements for placing diverting
material based on a state-of-the-art application of the technology.
The model used for all geometry estimates is a 2-D solution to the PKN model with
Carter's fluid-loss estimates. All other estimates of fracture width and fluid leakoff
after shut-in are from Nolte and Castillo. The PKN model is used because of the
manner in which the net pressure increases with injected volume for a confined-
height vertical fracture. The increase in pressure is treated as the major cause of
unwanted vertical height growth in the fracture. The assumption is that the in-situ
stress level of the adjacent zone will restrict fracture width-development. This
restriction will act as a bridging point for the physical barrier being created with the
diverting material.
The following job design parameters are used in the BRACKETFRAC computer
simulation.
Diverter Penetration  The BRACKETFRAC module allows the user to input the
desired length of projection or will calculate the required projection length.
• Input Projection Length The penetration required is evaluated with regard to
the feasibility of attempting a diversion treatment. The program ends if the
diverter penetration is less than 100 ft and reports these findings to the user. If
the desired diverter penetration is greater than 85% of the total fracture length,
this is also reported as an unrealistic option and the simulation ends.
• Calculated Projection Length  A prediction of net pressure at the end of the
fracture treatment is made by simulating the creation of the final geometry using
the pad fluid and maximum pump rate. This estimate of the net pressure which
would be generated during the final fracture treatment makes it possible to
determine at what distance from the wellbore the pressure becomes greater than
maximum allowable. This distance from the wellbore is selected as the
penetration needed for coverage by the diverting agent.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Fracture-Height-Containment Services
Page 21 of 22

Allowable Width  A minimum allowable net pressure is calculated if adequate


width can be generated at the maximum allowable net pressure. This minimum
allowable net pressure is defined as the pressure required to meet width
requirements relative to the diverting agent particle size and is used as the minimum
net pressure required at the end of injection in placement estimates.
Particle Rising and Settling Velocities  The estimates of rising and settling
velocities for the particles are calculated from the modified Swanson's equation with
a correction factor for clustered settling.
Diverting Treatment Estimate  The program maximizes rate and minimizes
viscosity in conjunction with the following conditions.
• Net pressure generated during the treatment is less than maximum allowable net
pressure.
• Fracture half-length and diverter placement is equal to the penetration required
for pressure diversion.
• Total proppant banking must not occur during slurry injection.
Staging  The diverting fracture will be created in a single stage if all placement
criteria are satisfied. The end of a stage of a diverting treatment can occur from the
following.
• Inadequate Fracture Width
• Excessive Net Pressure
• Proppant Banking
The program will attempt to simulate the placement of the diverter agents to the
desired depth of penetration in a maximum of three stages if single-stage placement
is impossible. For the second and third stages, the pressure limitation of staying
below maximum allowable net pressure is not evaluated at the wellbore. The point
of critical pressure is defined as the point in the fracture that was covered by
diverters in the previous stage.
Fluid Loss  Diverting agents must settle or rise to the desired position in the
fracture after the slurry has been placed in the fracture. The minimum distance the
diverting materials must rise or fall is the perforation height. The time required for
the diverting agents to “clear the perforations” is determined by calculating the
velocity of the slowest moving particle (for example, a J423 particle in an
INVERTAFRAC or BRACKETFRAC treatment).
The fracture must maintain adequate width while the particles migrate. If the fluid
leaks off faster than particle migration time, a flush stage equal to the difference in
time will have to be used. The flush stage neither increases the fracture volume nor
displaces the diverting particles in the horizontal direction.

DOWELL CONFIDENTIAL
Section 700.5
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Fracture-Height-Containment Services Dowell
Page 22 of 22

Diverter Scheduling  The artificial barrier created by the diverting materials is


assumed to be 12% of the gross height. The diverter is not scheduled in incremental
stages. For nominal diverting treatment design, the incremental changes in
concentration would be so small that staging would be difficult to implement
operationally. Therefore, diverters are mixed and pumped in a single concentration.

4.1 Additional Computer-Aided Job Design Information


The FracCADE Users Manual, and FracCADE Release Notes provide additional
computer-aided job design information.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 1 of 22

BREAKDOWN TECHNIQUES

1 Introductory Summary............................................................................................................. 2

2 Applications ............................................................................................................................. 3

3 Treatment Design..................................................................................................................... 3
3.1 Fracture Gradient................................................................................................................. 5
3.2 Fluid Selection ..................................................................................................................... 5
3.2.1 Solvent Selection Guidelines ...................................................................................... 6
3.3 Types of Formation Damage ............................................................................................... 8
3.4 Fluid-Loss Control.............................................................................................................. 13
3.5 Fluid Volumes .................................................................................................................... 13
3.6 Fluid Diversion ................................................................................................................... 14
3.6.1 Ball Sealers............................................................................................................... 14
3.6.2 Chemical Diverting Agents ....................................................................................... 15

4 Execution................................................................................................................................ 16

5 Evaluation............................................................................................................................... 17

6 Examples ................................................................................................................................ 18
6.1 Example No. 1 ................................................................................................................... 18
6.2 Example No. 2 ................................................................................................................... 20
FIGURES
Fig. 1. Ball catcher. .................................................................................................................... 17
Fig. 2. Friction loss of 15% HCl in 2.875 in. tubing. ................................................................... 22
TABLES
Table 1. Breakdown Design Methodology ................................................................................... 4
Table 2. The Dowell Solvent Formulations Most Commonly Used for Breakdown Fluids (The
Stimulation Materials  Acidizing Manual provides information for these systems)...... 7
Table 3. Summary of Formation Damage.................................................................................... 9

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 2 of 22

1 Introductory Summary
The purpose of a breakdown treatment is to establish positive communication
between the wellbore and reservoir (ensure that all perforations are open). This
manual section discusses the reasons for breakdown treatments and provides
guidelines for designing treatments. A breakdown treatment can be independent
from a later-planned fracture treatment or be included with it as a spearhead.
Specific well conditions will dictate the timing for the breakdown treatment.
Wells are often “broken down” after perforating and prior to pumping the fracturing
treatment. These treatments are often performed several days before the actual
fracturing treatment to allow formation testing. Breakdown treatments are usually
performed in cased wells to ensure that the casing has been effectively perforated
and the perforations are open. It is common for the perforation tunnel to be
extensively damaged, especially when wells are perforated in an overbalanced
condition (that is, the hydrostatic pressure of the well is greater than the reservoir
pressure). This damage is often severe enough to prevent the flow of stimulation
fluids from adequately entering the formation. A proper breakdown treatment
effectively removes perforation debris and eliminates most of the damage. At the
same time, this small treatment can confirm the validity of fracturing treatment
pressure calculations before an entire stimulation treatment crew is mobilized to the
wellsite.
Performing a breakdown treatment is very important when the interval has a limited
number of perforations or when the perforated interval exceeds 100 ft. If the well
contains a limited number of perforations and a breakdown has not been performed,
excessive treating pressures may be encountered because of an insufficient number
of perforations accepting fluid. When a large interval has been perforated, some of
the perforations may not accept fluid and that part of the interval may not be
adequately treated.
However, when a relatively small interval has been shot with a high perforation
density (for example, four shots per foot over approximately 25 ft), there may be no
advantage in performing a separate breakdown treatment. In this case, sufficient
perforations are present so that a fracture treatment can be started without
experiencing excessive pressures.
A breakdown treatment fractures the reservoir rock using a small amount of fluid. It
is normally performed in reservoirs where low permeability prevents matrix pump
rates. The fracture generated by a breakdown treatment is generally considered to
have a short half-length (tens of feet). Breakdown treatments are performed in both
sandstone and carbonate formations — usually those that have a permeability of
less than 10 md. The fluid used for these treatments should be compatible with
reservoir rock and wellbore fluids and be solids-free. It can be either reactive or
nonreactive, although a reactive fluid is normally more efficient. Fluid diversion may
be required and can be accomplished by any of several different plugging methods,
depending upon well and reservoir conditions. Pump rates are generally low and are
usually governed primarily by pressure limitations.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 3 of 22

An energized fluid may be used in a well with low bottomhole producing pressure to
promote cleanup after a breakdown treatment. Surfactants (foaming agents)
enhance gas entrainment and lower surface tension in the returning fluid and may
also be used to promote cleanup. The use of alcohol is also beneficial when the
bottomhole producing pressure is low. Alcohol decreases the vapor pressure,
surface tension and hydrostatic pressure, therefore enhancing cleanup.

2 Applications
The basic purpose for a breakdown treatment is to ensure that all perforations are
open. Several additional reasons for performing breakdown treatments follow.
• to establish a pump rate prior to a squeeze cementing operations
• to serve as a test of the cement sheath prior to stimulation work
• to determine the fracturing pressure gradient
• to lower the initial pressure of a future fracturing treatment
• to establish a channel through skin damage
• to open clogged perforations
• to assure adequate or selective fluid entry or both for subsequent pumping
procedures such as a limited entry treatment.

3 Treatment Design
A breakdown treatment exceeds fracture initiation pressure and creates a short
fracture. Fracture extension of only a few feet is expected. There will be no
increase in fracture conductivity because no proppant is introduced. Therefore, any
increase in production resulting from a breakdown treatment will be the result of
having removed existing skin damage.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 4 of 22

Refer to Table 1 for guidelines for breakdown treatment design.


Table 1. Breakdown Design Methodology
Step 1  Collect data.
• Well tests
• Reservoir tests
• Drilling/production records
• Laboratory test results of well and/or reservoir samples
Step 2  Is treatment required?
• Yes→Proceed to Step 3.
• No→No further action required.
Step 3  Is wellbore treatment needed before formation entry?
• Yes→Proceed to Step 4.
• No→Go to Step 5.
Step 4  Plan and execute a wellbore treatment (for example, acid cleaning,
corrosion inhibitor application or paraffin removal).
Step 5  What type of formation entry treatment is necessary and/or
acceptable?
1. Breakdown
• Yes→See Step 6.
• No→See Step 5-2 or take no further action.
2. Other
• Re-perforate, ABRASIJET* Service or other special procedure. Proceed to Step
6 or take no further action.
Step 6  Select breakdown fluid(s) and additive(s).
Step 7  Select pump rate not to exceed the maximum pressure limit.
Step 8  Determine if fluid diversion is necessary.
Yes→ Select method of diversion.
a) Ball sealers. Density? Number?
b) Temporary diverting agent. Liquid? Slurry? Stages? Amount per stage?
c) Permanent diversion. Cement? Liquid Chemical?
d) Mechanical downhole tool diversion.
No→Proceed to Step 9.
Step 9  Finalize the pumping schedule.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 5 of 22

3.1 Fracture Gradient


A breakdown treatment is a convenient time to ensure that the fracture gradient used
in the hydraulic fracture design is accurate. If the estimation is in error, the fracture
treatment can be revised. The fracture gradient can be estimated by taking an initial
shut-in pressure and applying Eq. 1.
pisi + ph
p fg = (1)
D
Where:
pfg = fracture gradient (psi/ft)
pisi = initial shut-in pressure (psi)
ph = hydrostatic pressure (psi)
D = depth to mid-perforation (ft).

3.2 Fluid Selection


Even though production will not be significantly affected by near-wellbore damage
once a propped fracture has been placed, a nondamaging fluid should be chosen as
a breakdown fluid. The fluid design should take the following parameters into
consideration.
• creation of emulsions
• creation of water-blocks
• rock wettability
• clay destabilization.
Any time water and oil are mixed together the possibility of creating an emulsion
exists. Special nonemulsifying surfactants can usually eliminate this problem.
Pretreatment compatibility testing between well fluids and breakdown fluids should
be conducted to identify the most effective nonemulsifier. This type of testing will
also reveal any tendency for the well fluids to drop out paraffins or asphaltenes.
Emulsions are not a problem in dry gas wells. However, these formations often tend
to retain water because of relative permeability and capillary pressure effects.
Breakdown fluids used in these formations should incorporate surfactants that are
efficient in lowering the interfacial tension properties of the fluid. To further minimize
the potential for a water-block, breakdown fluids in tight gas reservoirs may benefit
from pumping an energized fluid to reduce the amount of water put into the
formation. Nitrogen or carbon dioxide may be used as the gas phase. The fluid
phase can be a weak acid, water or a water and methanol blend.
Clay migration or swelling problems can be minimized by using a clay stabilizer in
addition to 2% (BWOW) potassium chloride water. Energized or foamed fluids may
also help minimize the effects of water on the clay particles.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 6 of 22

Dilute hydrochloric acids (7.5% or less) have long been accepted as common
breakdown fluids. These fluids are most appropriate in carbonate reservoirs where
the acid will react with the formation. In sandstone formations, the acid may not
have any material with which to react. A blend of 2% KCl water and methanol
incorporating a surfactant is, in some cases, as effective as acid. In sandstone
formations having carbonaceous cement, an acidic fluid may actually destabilize the
formation and allow the perforations to collapse. If acid is used as the breakdown
fluid, reducing agents or chelants or both should be added to the fluid to prevent iron
precipitation damage.
When damage to a well has been determined, identifying the type of damage is
necessary so that remedial action can be planned. Various types of damage can
exist because almost every operation performed on the well (drilling, completion,
workover, production and stimulation) is a potential source of damage. The physical
characteristics of the damage need to be known because they determine the desired
properties of the treating fluid.
The best aid in determining the cause of the problem is to obtain samples from the
well and have them analyzed. These can include formation core samples, sidewall
cores, water, oil, drilling mud, scale, paraffin and bacteria slimes. In short, any
material that may be considered as indicative of damage should be analyzed.
Pressure transient analysis can determine the extent of the damage.

3.2.1 Solvent Selection Guidelines


Formation breakdown can be achieved using any of the fluids summarized in
Table 2. Solvent selection will depend upon the nature of the material restricting
production. An acid-soak is advised in some cases before a breakdown treatment.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 7 of 22

Table 2. The Dowell Solvent Formulations Most Commonly Used for Breakdown Fluids
(The Stimulation Materials  Acidizing Manual provides information for these systems)
System Description Features and Benefits
Dowell X Inhibited 5% to 15% HCl Economical. Easily modified to suit a
variety of conditions from 80° to 400°F
(27° to 204°C).
Dowell Mud Acid Inhibited mixtures of HCl and Solvent for most siliceous minerals. Used
HF with Dowell X to dissolve damage from
mud and cement invasion as well as fines
and clays.
Alcoholic HCl+ or HCl or HCl + HF and F3 or Maintains low surface tension even in
Alcoholic Mud Acid+ K46 spent acid. Increases relative permeability
to oil and gas.
Gas Well Acid+ or Gas Mixtures of inhibited HCl and Used where low reservoir pressures
Well Mud Acid+ HCl + HF in alcohol retard cleanup. Good in dry gas wells.
BDA* Inhibited 5% to 15% HCl and Removes formation damage and reduces
special surfactants pumping pressures for subsequent
treatments. Disperses mud filter cake.
MSR 100, MSR 150 or Inhibited acid formulations with Exceptional silt and mud removal.
MSR 123 special dispersing and Removes mud damage. Minimizes
chelating agents. precipitation of dissolved iron.
MSR 123 contains HF and mutual
solvent. Used for formation cleaning.
Dowell Stabilized Acid Inhibited HCl with chelating Prevents precipitation of dissolved iron.
agents. Used in high iron content formations and
where iron scale is a problem.
Dowell Acid Dispersion Aromatic solvent in inhibited Hydrocarbons and acid soluble minerals
(DAD)+ HCl. that restrict production. Iron chelating
agents can be added.
NARS 200 and NARS Nonacid solutions of chelants For water- or acid-sensitive sandstone
201 and clay suspending agents. and carbonate formations. Works best at
temperatures above 150°F (66°C). Used
for either wellbore cleanup or breakdown
fluid.
+ These fluids contain flammable components that are classified as “high risk” (as defined in the
Dowell Location Safety Standards handbook). The Dowell Location Safety Standards handbook
provides approved handling methods for high risk fluids.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 8 of 22

Acid must contact the minerals that are to be dissolved. Oily, paraffinic coatings on
the surface of an acid soluble mineral will inhibit acid reaction. In such cases, use
appropriate surfactants, mutual solvents or aromatic solvent with the acid.
Acid-base fluids must be inhibited. Use the appropriate acid corrosion inhibitor,
giving consideration to temperature and exposure time of the pipe to the acid.
If the well is to be evaluated after breakdown and prior to fracture stimulation, it is
good practice to flow-back or swab treating fluids from the well as soon as possible
after breakdown. This will help assure that minimal damage will occur because of
fines migration in the reservoir rock or undesirable by-product deposition of
breakdown fluids. Energized fluid, surfactants or alcohol should be considered to
enhance cleanup in those wells with low bottomhole producing pressure.
Sandstones
Hydrochloric acid (HCl) is often used as a preflush ahead of an acid solution that
contains hydrofluoric acid (HF) for the following reasons:
• To minimize contact between HF and any wellbore or formation fluid that contains
potassium (K+), sodium (Na+) or calcium (Ca++) ions. This helps assure that
undesirable by-products caused by reaction of HF with these ions will be
reduced.
• To dissolve a maximum amount of carbonate (CO3- -) in order to minimize calcium
fluoride (CaF2)precipitation.
• To provide a low pH environment which will additionally hinder undesirable by-
product deposition from HF or spent HF acid-blends.
• Acetic acid can replace HCl in high temperature wells above 400°F (204°C).
Carbonates
HCl is often used as a stimulation fluid for the following reasons.
• The reaction between HCl and HCl-blends with carbonate minerals is extremely
fast when compared to the reaction rates of HCl with other naturally occurring
minerals or induced mineral scales. There is no advantage in using a more
expensive acid blend because the acid ions will preferentially spend on the
carbonate ions first.
• HCl will dissolve the formation away from any type of damage that may be
present providing the acid can be placed in contact with the formation. In this
case, positive communication with the reservoir can be attained, not by dissolving
damage in the formation, but by dissolving formation away from the damage.

3.3 Types of Formation Damage


Table 3 provides a summary of formation damage. Additional materials information
is provided in the Matrix Materials Manual and the Fracturing Materials Manual 
Additives.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 9 of 22

Table 3. Summary of Formation Damage


General
In those cases when the wellbore and perforations are clean
and the primary objective is to simply gain reservoir entry, the
fluid used for breakdown can be nonreactive but should be
tested prior to use for compatibility with well fluids and
reservoir rock.
Silts and Clays Effective Materials
Silt is composed of large clay-type minerals such as feldspar, • HCl
mica and chert. Clays are aluminosilicate minerals less than 4
• Mud acid
microns in diameter. Clay and silt together are called fines.
Damage due to silts and clays includes the invasion of the • MSR* acid
reservoir permeability by drilling mud filtrate or cement filtrate
(or both) causing the swelling or migration (or both) of • MSR water
reservoir formation fines. The HCl/HF solubility minus the HCl • Alcoholic acid
solubility is the total fines content. In addition, solids from the
drilling, completion or workover fluids can invade the formation • Alcoholic mud acid
and reduce permeability by migrating and plugging pore • Gas well acid
throats.
• Gas well mud acid
Filtrates and solids from mud and cement characterize the • BDA
damage encountered when performing a breakdown prior to a
remedial squeeze cementing operation. HCl or Mud Acid work • Stabilized acid
well in this application. • Nonaqueous acetic acid
• Aqueous acetic acid
• Organic mud acid
• Dowell acid dispersion (DAD)
• Formation cleaners
Wettability Change Effective Materials
Oil Wetting: Totally oil wetting a formation reduces the relative • Mutual solvents
permeability to hydrocarbons. This may occur due to
• Formation cleaners
adsorption of surface-active materials from oil-base drilling,
workover or completion fluids or from the native hydrocarbon. • Surfactants
Damage from oil-base drilling mud can fit into this category
(but can also cause emulsions). Wettability damage is
removed by the injection of mutual solvents to remove the oil
wetting hydrocarbon phase. In the case of native damage, this
can be followed by injection of strong water-wetting
surfactants to prolong the effect of treatment.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 10 of 22

Table 3. Summary of Formation Damage (continued)


Water Block (Water-wetting): A water block caused by an • Mutual solvents
increase in the water saturation near the wellbore decreases
• Formation cleaners
the relative permeability to hydrocarbons. A water block can
form either during drilling and completion operations through • Nonaqueous acid
invasion of a water-base filtrate or during production caused
by fingering or coning of formation water. The forming of a • Gas well acid
water block is enhanced by the presence of clays lining the • Alcoholic acid
pores of the reservoir rock. The hair-like structure of common
illite clay increases the adsorption of water on the pore walls. • Surfactants
• Alcohol
A water block is usually treated by reducing the interfacial
tension between water and oil (mutual solvents, surfactants).
Nonaqueous acids (alcoholic acids, nonaqueous acetic acid)
are particularly suitable in (gas) wells where water blocks exist
because they have solvent action on damaging solids,
increase the vapor pressure and reduce the surface tension
between water and gas.
Emulsions Effective Materials
The intermixing of oil- and water-base fluids in the near • Mutual Solvents
wellbore critical area or in the wellbore often results in the
• CLEAN SWEEP 1, 2, 3
creation of emulsions. Emulsions have high viscosity,
particularly water-in-oil emulsions. Typically, they are formed • Surfactants
due to invasion of drilling fluid filtrates, completion fluid filtrates
or treatment fluids into the reservoir rock. High pH filtrates
from mud or cement slurry or low pH filtrate from acidizing can
emulsify native oils. Similarly, hydrocarbon filtrates from oil-
base drilling or stimulation fluid can emulsify with formation
waters.
Emulsions can be stabilized by naturally occurring surface-
active agents (surfactants) found in formation oils or by
induced surfactants used in “treating” fluids. There are other
means of stabilizing emulsions, including fines present in live
or spent treating fluids and physical interaction along the fluid-
rock interface.
Generally, mutual solvent solutions, with or without a
demulsifier are used as remedial treating fluids for emulsions.
The specific carrier fluid for the mutual solvent or demulsifier
(or both) depends on the type of emulsion present (water-in-oil
or oil-in-water).
Scales
Scales are precipitated mineral deposits. They can precipitate
in the tubing, in perforations or in the formation.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 11 of 22

Table 3. Summary of Formation Damage (continued)


Scale deposition occurs, in most cases, as a result of
producing formation fluids. Lower temperatures or pressures
encountered in or near the wellbore disturb the chemical
equilibrium of formation water. They can also be deposited
from the interaction of formation water and introduced water.
• Carbonate scales include calcium carbonate (CaCO3), iron Effective Materials
(II) carbonate (FeCO3), strontium carbonate (SrCO3) or
• HCl
barium carbonate (BaCO3). CaCO3 is the most common
scale. It occurs in reservoirs rich in calcium and carbonate • Aqueous acetic acid
or bicarbonate ions. HCl will readily dissolve all carbonate
• For FeCO3, use HCl with iron
scales including SrCO3 and BaCO3. Simultaneous removal
stabilizers
of carbonate scale and oily, paraffinic deposits is
sometimes required. • Scale inhibitors/aids
• For oily carbonates, use Dowell
Acid Dispersion (DAD)
• Sulfate scales occur mainly as gypsum (CaSO4•2H2O) or
anhydrite (CaSO4). These scales are soluble in GYPBAN*
R1.
The less common strontium sulfate (SrSO4) is more difficult
to dissolve while barium sulfate (BaSO4) is insoluble in any
known solvent. It is necessary to mechanically remove
pure barium sulfate scales. However, these minerals are
frequently intermixed with gypsum/anhydrite and
carbonate scales, and as such the scale can be partially
removed by chemical solvent. These less soluble mineral
scales might then be more readily removed by mechanical
methods.
To dissolve gypsum, anhydrite or strontium sulfate with Effective Materials
GYPBAN R1, long contact times and higher temperatures
• GYPBAN R1
are beneficial. A 24 hr soak at temperatures above 150°F
(66°C) will dissolve most of these scales. Movement or • NARS 200, NARS 201
agitation of the solvent during the soak time will be
beneficial. • Hot HCl

Gypsum and anhydrite are also slowly soluble in hot


(212°F [100°C]) HCl.
• Chloride (salt) scales such as sodium chloride (NaCl) • Dilute solutions of HCl, KCl or
deposits are easily dissolved in fresh water or dilute NH4Cl
electrolyte solutions like KCl, NH4Cl or HCl. If fresh water is
to be used, care should be taken to prevent the fresh
water from contacting a water-sensitive reservoir and
causing damage.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 12 of 22

Table 3. Summary of Formation Damage (continued)


• Iron scales such as ferric sulfide (FeS) or iron oxide • FeS: HCl and iron stabilizers
(Fe2O3) are soluble in HCl. HCl with iron stabilizers, will
• Fe2O3, FeCO3: HCl and iron
dissolve these scales and prevent the precipitation of by-
stabilizers
products of the dissolved iron scales.
• Oily iron scale: HCl and/or iron
stabilizers and surfactants or
Dowell Acid Dispersion (DAD)
Iron scales are often formed as a result of corrosion to the
steel in the well and are deposited on downhole • Pickle Acid: HCl and iron
equipment. In these cases, the wellbore should be stabilizers
“pickled” with HCl containing iron stabilizers. After pickling,
the pickle acid should be circulated out of the well prior to
using any procedure that would cause the pickle acid to
invade the critical area of the reservoir adjacent to the
wellbore. Spent pickle acid could leave undesirable iron
hydroxide precipitate.
• Silica scales, when present, usually occur as very fine • Mud Acid or other acid
crystalline mineral deposits of opal or chalcedony. containing HF
Hydrofluoric acid can dissolve silica scales.
• Hydroxide or lime scales of magnesium or calcium • HCl
(Mg[OH]2 or Ca[OH] 2) will dissolve in any acid that will
sufficiently lower the pH of the scale environment and not
precipitate these chemical salts.
Organic Deposits
Organic deposits are divided into three categories. (1)
paraffins and asphaltenes precipitated from oil, (2) mixed
deposits, and (3) bacterial deposits.
• Paraffins and Asphaltenes Precipitated heavy Effective Materials
hydrocarbons and are typically deposited in the tubing,
• Paraffin/asphaltene dispersants
perforations or the formation. These deposits form
because of a change in temperature and pressure in or • Formation cleaners
near the wellbore during production. The heavy
hydrocarbon components of the formation oil do not • Mutual solvents
remain soluble in the oil and begin to crystallize in the • Toluene/surfactant blends
wellbore and in the critical area near the wellbore.
Cooldown of the wellbore or the injection of cold treating
fluids has the same overall effect as a temperature drop
during production.
Paraffin and asphaltene deposits can usually be dissolved
in organic solvents. Blends of solvents can be tailored to a
particular problem. Aromatics (toluene or xylene) are
efficient, general purpose solvents for paraffin or
asphaltene deposits. Addition of specific dispersants or

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 13 of 22

alcohols may make these solvents more efficient.


Table 3. Summary of Formation Damage (continued)
• Mixed Deposits Mixed-deposits can be blends of paraffin • Dowell Acid Dispersion (DAD)
or asphaltenes (or both) and either scales, silts or clays.
• Formation cleaners
Tiny crystals of precipitated heavy hydrocarbons can serve
as a nucleus for deposition of scales. Formation fines can
form nuclei in the same way. Mixed deposits require the
use of mixed solvents.
• Bacterial Deposits Bacteria are microorganisms that • Formation cleaners
multiply very rapidly in water under a broad range of
conditions. They can cause plugging at injection wellbores
in the critical near wellbore areas.
They can also cause severe corrosion problems. Sulfate • Dowell Acid Dispersion (DAD)
reducing bacteria generate hydrogen sulfide (H2S) which followed by formation cleaner or
attacks steel and causes pitting. The reaction produces biocides.
FeS scale. To alleviate this problem, the FeS scale must
first be dissolved to expose the bacteria to an effective
biocide.

3.4 Fluid-Loss Control


A fluid-loss additive is usually not required in the breakdown fluid, in fact, fluid loss is
desirable. A fluid capable of invading a few inches into the reservoir is ideal.

3.5 Fluid Volumes


The breakdown fluid quantity may be determined from field experience. If in an area
where inadequate or no previous treating history exists, consideration should be
given to the following parameters.
• cased hole or openhole
• bottomhole static temperature
• gross height
• net height
• number, size and density of perforations
• rock solubility
• the location of any deposits or damage (tubulars and perforations are critical
areas)
• the solubility of any deposits or damage
• should the wellbore be cleaned or otherwise treated prior to breakdown and
formation entry
• if a solvent or reactive fluid is to be used, what concentration will be best

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 14 of 22

• what is the stability of any solvent, reactive fluid or additive at well conditions
• should a preflush or overflush (or both) be used  which fluid should be used
• acid contact time in the pipe.
The basic reason for performing a breakdown treatment is to ensure that all
perforations are open. The fluid volume should be sufficient to gain formation entry
throughout the vertical extent of perforated interval. Diverting agents may be
required. The obvious purpose of perforations in the producing interval is to produce
reservoir fluids, hence all perforations should be open.
Most successful breakdown treatments typically use 50 to 400 gal of breakdown fluid
per foot of perforations depending upon the type, solubility and lateral extent of
damage present. Although as little as 10 gal of breakdown fluid per foot of
perforations have been used, such volumes are generally not successful in attaining
formation entry for the full vertical extent of the perforated interval. Volumes less
than 20 gal/ft are not recommended. Larger volumes (greater than 400 gal/ft) may
be required in special applications where diversion is difficult.

3.6 Fluid Diversion


Several fluid diversion methods may be used. These are:
• ball sealers
• chemical diverting (bridging) agents
• mechanical devices (for example, packer and bridge plug).
The volume of breakdown fluid may be dependent on the fluid diversion method.
Diverting agents may require larger carrier fluid volumes. Mechanical isolation using
a packer or bridge plug can either increase or decrease fluid volume.
Additional information on diverting techniques and materials is provided in Diverting
Techniques.

3.6.1 Ball Sealers


Ball sealers are the most commonly used diverting agent when a competent cement
sheath is present. It is common practice to use at least 50 to 100% more ball
sealers than there are perforations. The excess balls must be pumped because of
ball seating inefficiencies and placement timing. Breakdown treatments usually
consist of only a few thousand gallons of fluid, and all the ball sealers must therefore
be injected in a short period of time. Some ball sealers may be rendered ineffective
because they are already below a perforation when it opens and begins accepting
fluid. Ball sealers are most effective when there are no more than two perforations
per foot. Ball sealers can be used effectively at 1 ball every 20 gal of breakdown
fluid depending upon fluid pump rate and ball sealer density.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 15 of 22

Ball seating efficiency is dependent upon:


• fluid viscosity
• fluid velocity in the pipe
• fluid velocity through each perforation
• perforation size
• perforation shape
• density difference of ball sealer and fluid
• ball sealer size
• the quantity of ball sealers injected
• injection frequency (injected individually or slugged).
Ball sealers having a higher density than the fluid with which they are pumped have
a lower seating efficiency, but will fall into the rathole after the treatment and will
therefore not affect subsequent treatments.
Ball sealers having a lower density than the fluid with which they are pumped are
termed “buoyant ball sealers.” These ball sealers approach 100% seating efficiency
and are effective even at matrix rates, but they must be circulated out of the wellbore
before the fracturing treatment. If a buoyant ball sealer is left in the wellbore, it can
float in the wellbore fluids and reseat once the fracturing treatment begins. The
processes in which the buoyant ball sealers are used are protected by an Exxon
patent, therefore, Dowell is required to charge a license fee when buoyant ball
sealers are used.
Ball sealers are available in different sizes. A selection guide for sizing ball sealers
is provided in Diverting Techniques.
A good practice is to run tubulars or a junk basket past the entire perforated interval
to physically remove any remaining ball sealers before fracturing operations.

3.6.2 Chemical Diverting Agents


When the perforation density exceeds two shots per foot or where there may be
communication behind the pipe, chemical diverting agents are often the best choice
to gain entry into multiple intervals. These diverting agents can be continuously
added in small amounts to the breakdown fluid or slugged in concentrations of 50 to
500 lbm/1000 gal to achieve multiple zone entry. Surface indication of diversion is
usually easier to determine when the breakdown fluid is slugged with diverting agent
rather than when slowly and continuously added to the breakdown fluid. Conversely,
pressure surges are minimized when diverter is slowly added to the treating fluid.
Chemical bridging agents or inflatable packers are used for diversion in openhole.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 16 of 22

4 Execution
The breakdown treatment should be performed as designed, adhering to the steps
and procedures outlined by Dowell personnel and agreed to by the client.
Specific client directives must be followed unless they conflict with Dowell region,
area or district safe procedures and policies.
The treatment must be carried out in accordance with procedures outlined in the
Dowell Location Safety Standards handbook and should be thoroughly planned so
that unexpected events will be kept to a minimum.
Use the checklists shown in the Dowell Location Safety Standards handbook and the
checklists and policies set forth in region, area and district operational procedures.
Each of these are aids to help perform a better, safer treatment. Listed below are
events or situations that should be considered when planning and performing a
breakdown treatment. Best instruction will come, however, not from a manual, but
from field experience and from attention to recommendations from experienced
engineers and client representatives. They may use procedures different from those
listed here and will probably get good results, since each breakdown treatment is
both an art and a science. Encourage good engineering practices.
Good Breakdown Practices
• Establish an injection rate (using a nonreactive fluid), before any corrosive fluid is
pumped into the well. If this is not possible, ensure that the corrosive fluid can be
reversed out.
• Initial formation entry should be made at the lowest possible pressure required for
injection.
• Never attempt to reverse-out with a ball injector in the treating line. There is
always a chance for a ball sealer to be accidentally injected into the annulus.
• Know when specific fluids, ball sealers or diverting agents are to reach the zone
to be treated.
• Normally, pressure response due to ball sealers should be expected near or just
over the wellbore displacement volume. A pressure increase should be expected
at any time regardless of spacing during injection.
• Diverting agents are designed to bridge in a fracture. Allow for at least 400 gal
over wellbore displacement for pressure increase due to diverting agent.
• Premature maximum pumping pressure may be reached during a breakdown
treatment when using ball sealers or diverting agents. Be prepared for flowback
procedures.
• If ball sealers are expected to flow back to the surface, a ball catcher (Fig. 1)
should be used to prevent plugging the choke assembly. The ball catcher should
be installed between the tee on the wellhead and the first valve. An additional
bleedoff tee may be installed on the ball catcher assembly to facilitate removal of

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 17 of 22

the ball sealers. Part numbers for ball catchers are provided in the treating
adapters section of the Treating Equipment Manual.
• The use of nitrogen or carbon dioxide can dampen the pressure increase
associated with ball sealers or diverting agents.
• A ball catcher is recommended when using ball sealers (of any density) with
nitrogen or carbon dioxide.
• Where there is a good cement sheath, ball injection at regular intervals should
provide good fluid diversion.
• Where a poor cement sheath is suspected, use of a diverting agent may be best.
• If the cement bond is poor, injection of ball sealers in slugs may give better
results than if injected individually. (Communication behind the pipe can
decrease seating efficiency.)
• Optimum ball sealer seating efficiency (conventional ball sealers) is obtained
when the minimum flow velocity inside the pipe is 150 ft/min and the minimum
flow velocity in the perforation tunnels is 750 ft/min.
• When treating down tubing, under a packer, the best seating efficiency will occur
if the packer is set no closer than 200 ft from the top perforation.

Fig. 1. Ball catcher.

5 Evaluation
A breakdown treatment is successful if wellbore-to-reservoir communication is
established.
There may be other questions.
• Did the cement sheath give a positive test?

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 18 of 22

• Was a good production test obtained?


• Was sufficient formation entry indicated to proceed with the next treatment?
Answers to these questions are dependent on a successful breakdown treatment.

6 Examples
There can be many specific designs appropriate to breaking down a well, depending
upon the experience of the individual designing the treatment and what works best in
a specific area.
Two examples are shown. In the first example, formation entry will be attempted.
The second example is a treatment using ball sealers to “open” perforations.
The designs in Section 6.1 and 6.2 are an appropriate design for stated conditions.
There may be other designs appropriate for this situation.

6.1 Example No. 1


Well and Reservoir Conditions
Casing : 5.5 in., 15.5 lbm/ft
Tubing : 2.875 in., 6.5 lbm/ft, H40
Packer : Retrievable at 5450 ft
Number of Perforations : 30
Average Diameter of Perforations : 0.4 in.
Perforated Interval : 5600 to 5650 ft
TVD to Mid-Perforation : 5625 ft
Well Type : Gas
Reservoir Type : Limestone
Gross height : 200 ft
Net height : 50 ft
Fracture gradient : 0.70 psi/ft
Bottomhole static temperature : 150°F (66°C)
Bottomhole producing pressure : 2500 psi
Treatment Parameters
The cement bond log (CBL) indicates a good bond across the reservoir. Prior results
in the same area have shown 15% HCl, plus additives, to be an appropriate fluid.
Maximum surface tubing pressure is 4200 psi. The internal yield pressure of the
tubing is 5283 psi (from the Dowell Field Data Handbook). Unless otherwise

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 19 of 22

specified by a company representative, use 80% of the internal yield pressure as the
maximum surface tubing pressure. Therefore, 80% of 5283 psi is approximately
4200 psi.
Design
1. Load the wellbore with 2% KCl water.
2. Circulate 250 gal of 15% HCl, plus additives to 5400 ft.
3. Close the packer and pressure the annulus as necessary.
4. Start pumping at low rate. Increase pressure until the fracture initiates. Increase
the pump rate to ensure fracture propagation.
5. Flush with 2% KCl water.
Execute
The surface pressure was gradually increased until a pressure break occurred. A
constant rate of 4 bbl/min at 2050 psi was obtained. An initial shut-in pressure of
1475 psi was recorded.
Evaluate
1. The pump rate (4 bbl/min) and wellhead pressure (2050 psi) indicate that the
formation was broken down.
2. Calculate the fracture gradient from the treatment data (Eq. 1).
pisi + ph
p fg =
D
1475 + 2464
p fg = (1)
5625
p fg = 0.70

The known reservoir fracture gradient (0.70 psi/ft) and the calculated fracture
gradient (0.70 psi/ft) are the same. This, plus the observed pressure break indicate
that the formation was broken down.
3. Calculate the perforation friction loss (Eq. 2). Use Fig. 2 to determine the tubular
friction pressure.
p pf = pw − pisi − ptf (2)

Where:
ppf = perforation friction pressure (psi)
pw = surface pressure (psi)
pisi = initial shut-in pressure (psi)
ptf = tubular friction pressure (psi).

ppf = 2050 - 1475 - (90 psi/1000ft) (5625 ft)

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 20 of 22

ppf = 2050 - 1475 - 506


ppf = 69 psi
4. Calculate the number of open perforations (Eq. 3).

qi 2 p
n= (3)
p pf d pf 4 ( 0.323)

Where:
n = number of open perforations
qi = pump rate (bbl/min)
ρ= fluid density (SG)
ppf = perforation friction pressure (psi)
dpf = average diameter of perforations (in.).

n=
( 4) 2 ( 1)
(69)(.4) 4 ( 0.323)
n = 5 perforations open

The relatively few open perforations (17%) indicate that another breakdown
treatment should be performed using diverting agents (that is, ball sealers).

6.2 Example No. 2


Well and Reservoir Conditions
Same as Section 6.1.
Treatment Parameters
This well was previously acidized with 250 gal of 15% HCl plus additives.
Calculations indicate that 17% of the perforations are open.
Maximum surface tubing pressure is 4200 psi. The internal yield pressure of the
tubing is 5283 psi (from the Dowell Field Data Handbook). Unless otherwise
specified by a company representative, use 80% of the internal yield pressure as the
maximum surface tubing pressure. Therefore, 80% of 5283 psi is approximately
4200 psi.
Design
1. (Optional) Swab or flow as much liquid as possible out of the tubing. This will
prevent injection of large quantities of wellbore fluid into the reservoir.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Breakdown Techniques
Page 21 of 22

2. Select a pump rate that will propagate the fracture and provide optimum ball
sealer seating efficiency. At 5 bbl/min, fluid velocity is 864 ft/min in the tubing
and 1072 ft/min in the perforation tunnels, assuming all perforations are open.
Velocity (and friction) in the perforation tunnels will increase as the number of
open perforations decrease.
3. Close the packer and pressure the annulus as necessary.
4. Acidize using 2500 gal of 15% HCl plus additives and 60 (100% excess) ball
sealers (0.75 in. diameter, 1.3 SG). Pump 150 gal of acid and evenly space the
ball sealers in the remaining 2350 gal of acid. If maximum pressure is attained,
surge ball sealers off of the perforations by flowing back three to four barrels of
fluid. Suspend pumping operations for 1/2 to 1 hr to allow the ball sealers to
settle into the rathole.
5. Flush to the bottom perforation with 2% KCl water.
6. Swab or flow back the well as soon as possible (after Dowell has rigged down).
Execute
1. Calculate the wellhead treatment pressure at 5 bbl/min and negligible perforation
friction (Eq. 4). Use Fig. 2 to determine the tubular friction pressure.
Hydrostatic pressure = (0.465 psi/ft) (5625 ft) = 2616 psi
Tubular friction pressure = (140 psi/1000 ft) (5625) = 788 psi
Perforation friction pressure = 0 psi
Reservoir friction pressure = (0.7 psi/ft) (5625 ft) = 3938 psi
pw = piw + ptf + p pf − ph (4)

Where:
pw = surface pressure (psi)
piw = reservoir fracturing pressure (psi)
ptf = tubular friction pressure (psi)
ppf = perforation friction pressure (psi)
ph = hydrostatic pressure (psi).
pw=3938+788+0-2616
pw=2110 psi
2. Execute according to the design. Maximum pressure may be attained at any
time after the first ball sealer has reached the perforations.

DOWELL CONFIDENTIAL
Section 700.6
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Breakdown Techniques Dowell
Page 22 of 22

Evaluate
Calculate the number of open perforations using. If the calculations indicate that
most or all of the perforations are open, proceed with testing or further stimulation. If
the calculations indicate that few perforations are open, determine the next action
(that is, perforate again, acidize again).

Fig. 2. Friction loss of 15% HCl in 2.875 in. tubing.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 1 of 40

DIVERTING TECHNIQUES

1 Introductory Summary............................................................................................................. 2

2 Application ............................................................................................................................... 4
2.1 Ball Sealers .......................................................................................................................... 5
2.1.1 Buoyant Ball Sealers .................................................................................................. 5
2.1.2 Conventional Ball Sealers........................................................................................... 8
2.1.3 Temperature and Pressure Effects........................................................................... 13
2.1.4 Effects of Various Fluids ........................................................................................... 13
2.2 Limited Entry ...................................................................................................................... 14
2.2.1 Perforation Erosion ................................................................................................... 18
2.2.2 Contributing Factors ................................................................................................. 19
2.2.2.1 Casing Size.................................................................................................... 19
2.2.2.2 Number and Size of Perforations................................................................... 19
2.2.2.3 Differences in Fracturing Pressures............................................................... 21
2.2.2.4 Pressure Limitations ...................................................................................... 21
2.2.2.5 Hydraulic Horsepower Requirements ............................................................ 21
2.2.2.6 Perforation Friction Pressure ......................................................................... 21
2.2.2.7 Breakdown Technique ................................................................................... 22
2.2.2.8 Determination of Formation Fracturing Pressure........................................... 22
2.2.3 Design ...................................................................................................................... 22
2.3 Single-Point Entry .............................................................................................................. 33
2.4 Plugback (Pine Island) Technique ..................................................................................... 34
2.5 Slurried Solids .................................................................................................................... 36
2.6 Viscous Fluids .................................................................................................................... 37
2.7 Baffles and Balls ................................................................................................................ 38
2.8 Bridge Plugs....................................................................................................................... 40
FIGURES
Fig. 1. Rising velocity of 7/8-in. OD buoyant ball sealers............................................................. 7
Fig. 2. This graph provides information necessary for calculating the injection rate to
“balance” buoyant balls in different sizes of pipe. ............................................................. 7

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 2 of 40

Fig. 3. Example printout of the PERFBALL ball-sealer-performance computer software.............9


Fig. 4. This graph contrasts the pump rate versus fluid velocity in different sizes of pipes. .......12
Fig. 5. This graph contrasts the pump rate versus fluid velocity in different sizes of
perforations. ....................................................................................................................12
Fig. 6. Core selection guide for temperature and pressure. .......................................................13
Fig. 7. Injection rate increases thus friction pressure increases.................................................15
Fig. 8. Nomograph for determining friction through a perforation...............................................16
Fig. 9. Reference to determine the injection rate per perforation using fresh water...................17
Fig. 10. Higher density fluids exhibit higher perforation friction pressures. ................................18
Fig. 11. Change in perforation friction versus perforation coefficient. ........................................20
Fig. 12. Gamma ray-sonic log. ...................................................................................................23
Fig. 13. Friction of WF140 in 4 1/2 and 5 1/2 in. pipe.................................................................25
Fig. 14. Plugback Pine Island technique. ...................................................................................34
Fig. 15. Two-stage fracturing procedure, using baffle and ball technique. .................................39
TABLES
Table 1. Perforation Size versus Ball-Sealer Core Size ...............................................................5
Table 2. Example for Selecting a Perforation Schedule for Multiple Zones Having Different
Bottomhole Fracturing Pressures .................................................................................32
Table 3. Diverting Agents ...........................................................................................................36

1 Introductory Summary
The purpose of using a diverting agent or a diverting technique while hydraulic
fracturing is to assure that each separate formation interval is fractured. Commonly,
ball sealers, slurried diverting agents or limited entry treatments are used to
effectively create multiple fractures during one continuous fracturing operation. This
requires staged fracturing treatments, where each stage is identical and each zone
receives its own pad volume and graduated proppant schedule. At the end of each
stage, enough diverter is injected to either cover the perforations of the zone being
treated or bridge in the fracture of the zone being treated, effectively diverting the
next stage into other perforations. The diverter is then followed by the pad volume to
initiate the fracture in the next zone and the process is repeated; the diverter at the
end of the final stage is eliminated.
In thin reservoirs with only one zone, it is relatively easy to cover the entire vertical
interval with a single-stage treatment (assuming a vertical fracture). However, many
wells do not have a thin single zone. Some do not fracture vertically. In these wells,
diversion must be used to direct each separate stage of the treatment into each of
the different zones of the perforated interval. Well conditions which may require
diversion are
• multiple sets of perforations separated by blank casing
• thick, massive formations
• multiple zones or stringers
• small tubulars which necessitate lower injection rates

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 3 of 40

• horizontal fractures.
Fluids pumped at fracturing pressures will first enter and fracture the zones that have
lower fracturing pressures. It becomes necessary to temporarily isolate the zones
that have the lower pressures and divert the fracturing fluid injection to the zones
that have the higher fracturing pressures. Diversion can occur inside the pipe at the
perforations, in a channel in the casing/formation annulus or in the fracture itself.
Diversion techniques are:
• use of ball sealers
• limited entry
• single-point entry
• plugback
• use of slurried solids
• use of viscous fluids
• frac baffles
• packers and bridge plugs.
Combinations of these techniques can also be effective; the method to be used
depends on the experience in the area, well configuration and economics. The
fracture-design engineer must evaluate each method for a particular reservoir and
the resulting reservoir performance as a function of the treatment diversion
techniques to determine the optimum method for any particular situation.
The cost of fracture diversion should not be the sole indicator of the most economical
diversion technique. Use of the FracNPV* software as a function of the treatment
diversion technique should be studied. It is quite likely that the most expensive
diverting technique is also the most economical diverting technique, because it may
improve the cumulative production and cash flow.
Jobs using the ball sealer diversion, limited entry or modified limited entry diversion,
and diversion using slurried solids can all be performed continuously (without
stopping the pumping operations). Such treatments are performed with multiple sets
of exposed perforations. During treatment, there is no way to determine which of the
exposed zones or sets of perforations is accepting fluid. Proof of diversion into
multiple zones after a continuous fracturing operation is best determined by using
the appropriate wireline logging procedure, such as a temperature or a radioactive
survey.
The plugback technique, most operations using viscous fluids, operations using a
packer and bridge plug, and operations using a frac baffle are not considered true
continuous fracturing treatments. Shutdown for zone isolation is necessary when
using a non-continuous fracture treatment technique.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 4 of 40

2 Application
When completing a well having multiple intervals or one with a long interval to be
perforated, there are general factors to consider that can have an effect on the
fracture diversion.
A sound primary cement job is critical in well completion and stimulation. If the
cement does not provide isolation between sets of perforations, one zone will
fracture initially in a multistage treatment; subsequent stages will feed through
channels in the cement into the same fracture. Even worse, a fracture will occur at
some undesired zone outside the interval selected for treatment. Abnormal
production may indicate this condition; more often production is poor and, due to the
many uncertainties involved, the true reason for poor production may go
unrecognized. However, zone isolation of poorly cemented wells during fracturing
has been accomplished using slurried solids or viscous fluids.
The number and location of perforations in a casing string completion are important
to the extent that they can help or hinder the injection of fracturing fluids. If there are
too many, the flow rate of a moderate-to-low viscosity fluid may not be sufficient to
carry proppant into the perforation tunnel. If there are too few, the large pressure
drop across the perforations may unnecessarily increase hydraulic horsepower costs
and restrict overall injectivity. Perforation size and density should be carefully
engineered to not only satisfy anticipated production rates, but also to meet the
requirements of a fracture stimulation treatment.
General Rules Concerning Perforations and Fracture Diversion
• A large number of perforations are not normally required to produce at the top
allowable rate, nor for the best ultimate recovery.
• In a multistage, multizone fracturing treatment, ball sealers cannot effectively
divert a treatment stage unless the separate groups of perforations can be
isolated from the other groups. Thus, it is beneficial to know the number of
perforations accepting fluid per stage. In addition, there must be sufficient fluid
viscosity and differential pressure across the perforations to seat and seal the ball
sealers.
• It is widely understood that most perforations, even when they are shot under
ideal conditions, do not guarantee effective communication between the reservoir
and the wellbore. A “ball-out” treatment with a nondamaging fluid and ball
sealers is usually necessary to assure that all perforations are open.
• The most efficient use of hydraulic horsepower is to treat one zone at a time.
• Slurried solids can be very effective diverters, but in wells perforated with a
limited number of perforations, ball sealers are more reliable.
• Some fracturing fluids may suffer irreversible viscosity reduction caused by shear
degradation through the perforations, the higher shear rates occurring in smaller
diameter perforations. This factor should be considered when planning a

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 5 of 40

perforation schedule for a well to be fracture stimulated. Consult the Fracturing


Materials Manual  Fluids for more information about shear sensitivity.
• In openhole completions or in cased holes with many perforations, an appropriate
electric log survey after each stage is necessary to establish the vertical footage
which was treated by the preceding stage.

2.1 Ball Sealers


Ball sealers are small spheres used to seal perforations and divert treating fluid to
other perforations during breakdown and fracture stimulation. They are coated with
rubber to provide a better seal. Their core can be made of nylon, high-strength
phenolic, aluminum or glass microspheres bonded with epoxy (buoyant balls). Ball
size is dictated by the diameter of the perforation, the core should have a 25%
greater diameter than the perforation to avoid the risk of permanently seating the ball
in the perforation. The most common ball sealer size is one with a 3/4-in. core with a
1/8-in. coating of rubber. The OD of these balls is 7/8 in. Refer to Table 1.
Table 1. Perforation Size versus Ball-Sealer Core Size
Perforation Size (in.) Suggested Core Size (in.)
Less than 3/8 1/2
3/8 through 1/2 3/4
5/8 7/8
Larger than 5/8 1-1/8
Solid rubber balls should never be used, because they are relatively easy to extrude
through perforations.
There are two basic types of ball sealers; buoyant and conventional.

2.1.1 Buoyant Ball Sealers


A buoyant ball sealer is one that will float in its carrier fluid. Buoyant balls are very
efficient, nearly 100% seat on the perforations. They have an excellent application in
matrix treatment diversion and have been successfully used in fracture stimulation.
Nearly all buoyant ball sealers will be returned to the surface where special safety
equipment and procedures must be used to retrieve them.
Because the buoyant ball sealer and the carrier fluid will be moving at different rates,
it is often desirable to know where the ball sealer is located at any given point in time
in relation to the perforations. This can only be done by keeping an accurate log of
the fluid volumes pumped and time elapsed after injecting the balls. The calculations
involved may be better understood by following the example below. Fig. 1 is helpful
in solving this problem. Specific pipe and fluid data are provided in the Dowell Field
Data Handbook.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 6 of 40

Example  Buoyant Ball-Sealer Displacement


Given: 25 buoyant ball sealers, each having a specific gravity of 0.9, are to be used
in 15% HCl to break down the perforations from 5230 to 5236 ft. Treatment will be
down 5-1/2 in., in 15.5 lbm/ft casing at 5 bbl/min.
Determine: (1) the time required to seat the last ball, and (2) the displacement
volume required to seat the last buoyant ball.
Note that the pipe volume PLUS a volume of fluid to compensate for the rising
velocity of the balls must follow the last ball injected.
1. Casing volume to the bottom perforation is
(5236 ft)(0.0238 bbl/ft) = 124.6 bbl.
2. Time to pump the casing volume is
124.6 bbl
= 124.6 bbl
5 bbl / min
3. Density differential between the acid and the buoyant balls is
1.075 - 0.9 = 0.175.
4. The rising velocity of the buoyant balls in 15% HCl is 63 ft/min (Fig. 1).
5. Capacity of the casing is
42.0168 ft/bbl.
6. Therefore, the downward velocity of the buoyant ball sealers due to the fluid rate
of 5 bbl/min is
5 bbl/min x 42.0168 ft/bbl = 210 ft/min.
7. The resultant downward velocity of the balls at 5 bbl/min becomes
210 ft/min (downward) - 63 ft/min (upward) = 147 ft/min (downward).
8. The pumping time to seat the last ball is the perforation depth divided by the
resultant downward velocity of the ball
5236 ft
= 35.6 min
147 ft / min

9. Displacement volume required to seat the last ball is


5 bbl/min x 35.6 min = 178 bbl.
From these example calculations, (1) the time required to seat the last ball is
35.6 min after injection, and (2) the volume of the displacement fluid required to seat
the last buoyant ball sealer is 178 bbl.
Under specific conditions, the use of buoyant ball sealers requires the collection of a
license fee. These conditions are explained in the current Dowell Price Book.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 7 of 40

Fig. 1. Rising velocity of 7/8-in. OD buoyant ball sealers.

Fig. 2. This graph provides information necessary for calculating the injection rate to
“balance” buoyant balls in different sizes of pipe.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 8 of 40

2.1.2 Conventional Ball Sealers


Conventional ball sealers are those having a greater density than their carrier fluid.
Depending on the density needed, these ball sealers can be obtained with specific
gravities of from 0.9 to 1.4, in 0.1 increments. Conventional ball sealers are
satisfactorily and routinely used in all types of wells to divert the treating fluids from
one perforation to another. They are most effective when there are a limited number
of relatively isolated perforations. Accurate prediction of the ball action from well to
well depends on good cement bond, similar perforating schedules and similar
treatments. In most cases, the ball action in a given well is difficult to predict.
Depending on the pump rate, relative density differences between the ball sealer and
the carrier fluid, carrier fluid viscosity, perforation friction and other inertia and drag
factors, extra ball sealers are frequently used. Dowell uses the PERFBALL
computer software to predict ball sealer performance under dynamic conditions. An
example worksheet and printout of the calculations are shown in Fig. 3.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 9 of 40

Fig. 3. Example printout of the PERFBALL ball-sealer-performance computer software.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 10 of 40

A total of 10% to 100% excess ball sealers is commonly used, depending on the
differential pressure at the perforations. In most situations, conventional ball sealers
will fall to the bottom of the well after being dislodged from the perforations. At the
bottom, they are no longer available to move up the wellbore to plug or partially plug
the surface equipment. Return of the ball sealers to the surface, unless advance
preparations are made to handle them, can be a safety hazard. Buoyant ball sealers
can also interfere with swabbing operations.
In a fracturing treatment where two or more perforated intervals are exposed and
ball sealers are to be staged, the treating fluid is pumped into the well and enters the
perforations opposite the zone having the lowest fracturing pressure. Ball sealers
are injected in the first part of the flush fluid following the fracture treatment of the
first zone. Since this perforated interval readily takes fluid, the injected ball sealers
follow the fluid stream into the perforations where they seat. Differential pressure
holds them in place. While the pressure on the well is maintained, the second
fracturing stage immediately follows the first stage flush. Since the perforations
opposite the first zone are sealed, the pressure builds up and the formation opposite
the second perforated interval is broken down and the fracturing fluid is injected.
This process is repeated as many times as is necessary to complete the multistage
treatment. On completion of the fracturing treatment, the pressure on the wellhead
is released, and the differential pressure from the formation to the wellbore causes
the balls to be released from the perforations. The density of the conventional ball
sealers, being heavier than the fluid, causes them to fall to the bottom of the well.
The following parameters must be controlled to assure the best sealing efficiency of
the conventional ball sealers.
• flow velocity in the pipe and through the perforations
• relative size of ball sealer and perforation
• perforation shape
• relative density of ball sealer and carrier fluid
• viscosity of the carrier fluid
To be practical and effective in sealing off casing perforations in multistage fracturing
treatments, the ball sealers must meet the following conditions.
• They must be of appropriate size and density as to be retained in the fluid stream
so that the ball will be directed to and held on a perforation.
• Ball sealers must be tough enough not to be extruded through the perforation
under the pressure differentials actually encountered in the field. Ball sealers
extruded through perforations can plug the perforations from the reservoir side.
• They must not only adequately seal the perforation during treatment, but also
free themselves from the perforation when the pressure differential is decreased.
• They should be of appropriate density so they will not be a problem when
released from the perforation; they should either fall into the rathole or be quickly

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 11 of 40

returned to the surface and captured in a ball catcher. If the presence of ball
sealers at the bottom of the well is objectionable, they may be removed with a
bailer. In most cases, it is desirable to leave the balls on the bottom.
• If there is communication behind the casing due to a poor cement bond or due to
circulation behind the pipe between closely spaced perforations, ball sealers will
not provide the desired results.
• Flow-velocity guidelines for the best ball action are approximately 150 ft/min
inside the pipe and 750 ft/min through the individual perforation. The maximum
flow velocity inside the pipe for good ball action is 1000 ft/min (Fig. 4 and Fig. 5).
Fig. 4 shows the velocity of fluid in various sizes of pipe at different pump rates.
The fluid viscosity and density differential between the carrier fluid and the ball
sealer will also have an effect on the ball sealer velocity. Fig. 5 illustrates the
velocity in different sizes of perforations at different rates.
• When treating down the tubing under a packer, the best results are obtained if
the packer is set to no closer than 400 ft of the top perforation. Closer, the nozzle
effect and higher velocity of the fluid in the smaller diameter tubing can speed the
ball past the perforation. Perforations should be few, round and burrless. Ball
sealers should be only slightly more dense than the carrier fluid and should also
be more dense than the fluids to be returned from the well. Balls must have a
hard core; the diameter of the core should be larger than the perforation diameter
and the covering on the ball should be deformable to fit the perforation geometry.
• The best individual ball action will be obtained if the balls are spaced evenly in
the injection fluid, that is, one ball per 1/2 bbl of fluid or similar schedule. If there
are more than two perforations per foot or if communication behind the pipe is
suspected, injecting the balls in batches may be best. The batch injection of balls
may help assure that they all reach the perforations at the same time and may
seal several perforations simultaneously, thus reducing the chance of being
unseated when more balls arrive.
• The primary disadvantage of the ball sealer diversion during fracturing is that if
several intervals are simultaneously exposed, there is no assurance that each
zone is positively treated and excluded. A ball-out breakdown job with ball
sealers prior to the main fracture-stimulation procedure will increase the
likelihood of full zonal coverage.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 12 of 40

Fig. 4. This graph contrasts the pump rate versus fluid velocity in different sizes of
pipes.

Fig. 5. This graph contrasts the pump rate versus fluid velocity in different sizes of

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 13 of 40

perforations.

2.1.3 Temperature and Pressure Effects


Ultimately, it is differential pressure across the perforation that determines whether
or not a ball will seat. Practical experience has shown that a 75 to 150-psi pressure
drop across the perforation is required to seat and hold the ball sealers on the
perforations. Nylon cores will extrude excessively if exposed to high pressures and
temperatures. Use the phenolic cores when the fluid temperature is expected to
exceed 250°F (121°C) or 200°F (93°C) if the change in pressure will be greater than
5000 psi. However, above 200°F (121°C) and 5000-psi differential pressure, a ball
having harder rubber is required (like the phenolic core ball). Fig. 6 is an illustration
of the temperature and pressure constraints. The phenolic core balls have been
tested to 300°F (149°C) and a 10,000-psi differential and a striking velocity of 16,000
ft/min. Subjected to 300°F (149°C) and 9900 psi for 30 min, they released from the
test perforations at less than 25-psi backpressure.

2.1.4 Effects of Various Fluids


The standard covering of a nylon core ball is 1/8 in. of neoprene rubber. Neoprene
rubber allows sufficient deformation to seal in an average perforation without
rupturing. It is not affected by brine, hydrochloric acids or crude oils; however,
aromatic hydrocarbons (like toluene and xylene) and strong organic acids (like formic
acid) will cause the neoprene to swell and soften, even at low temperatures with no
pressure.
The ball with the phenolic core has a more fluid resistant cover.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 14 of 40

Fig. 6. Core selection guide for temperature and pressure.

2.2 Limited Entry


Limited entry uses a limited number of perforations to provide an intentionally high
pressure drop across the perforations at a designed treating rate. The high pressure
drop across the perforations is designed to be high enough to cause the bottomhole
treating pressure in the casing to exceed the various fracturing pressures of the
perforated zones.
Limited entry provides for the equal distribution of treating fluids through all
perforations by controlling the differential pressure across all perforations. Correctly
applied under the proper well conditions, limited entry can be very effective and can
result in considerable savings in well completion costs, particularly in deep wells.
The process is not difficult to apply.
For the limited entry technique to work, a good cement bond is needed around the
casing, the proper number and size of perforations must be placed in each porous
interval, and barriers to fracture growth must exist between each porous interval.
The limited entry technique should be used to pump into multiple zones where
multiple fractures will be created. If no barriers to fracture growth exist, and it is
expected that the fracture created by the treatment will be a single, continuous
vertical fracture, then it does not matter where the fracturing fluid leaves the wellbore
and enters the hydraulic fracture. Limited entry is to be used only where multiple
vertical fractures are going to be created.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 15 of 40

To obtain the best selectivity and the best control of the stimulation fluids with the
limited entry technique, a ball-out breakdown job should be performed prior to the
fracture stimulation. Ball-out should be performed in conjunction with using a packer
and bridge plug to selectively isolate each perforated zone, otherwise positive
assurance of all perforations being open is negated. A ball-out breakdown job prior
to the actual limited entry job is a must.
The key to limited entry is to use perforation friction to increase the pressure inside
the pipe to a pressure greater than that necessary to fracture treat all perforated
intervals. Fig. 7 illustrates that perforation friction varies directly with the rate of the
fluid being injected through the perforation. By controlling the injection rate per
perforation, the pressure inside the pipe is also controlled and can be increased to
exceed the fracturing pressure of all perforated intervals. Fig. 7 shows a 750-psi
friction pressure at 2.3 bbl/min through one 0.375-in. diameter perforation. Likewise,
a 750-psi friction pressure will be generated at a pump rate of 23 bbl/min through ten
0.375-in. diameter perforations; each perforation acts as a choke, with a differential
pressure of 750 psi across each perforation.

Fig. 7. Injection rate increases thus friction pressure increases.


Note: This graph shows that as the injection rate increases through a perforation,
the friction pressure also increases. The rates and pressures indicated here are
specific to the conditions shown.
Fluid density affects the perforation friction while fluid viscosity does not. Higher
density fluids exhibit higher perforation friction pressures (Fig. 10).

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 16 of 40

A nomograph (Fig. 8) was prepared using Eq. 2. The nomograph provides a graphic
method for solving the equation. Fig. 9 can be used as a quick reference to
determine the injection rate per perforation using fresh water.

Fig. 8. Nomograph for determining friction through a perforation.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 17 of 40

Fig. 9. Reference to determine the injection rate per perforation using fresh water.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 18 of 40

Fig. 10. Higher density fluids exhibit higher perforation friction pressures.

2.2.1 Perforation Erosion


In fracturing treatments using a proppant, the surface pumping pressures usually
decrease when the proppant is added because the hydrostatic pressure increases.
In limited entry treatments using a propping agent, the surface pressures also
decrease, not only because of an increase in the hydrostatic pressure due to the
proppant but also because erosion smoothes and enlarges the perforations,
reducing the friction pressure at the perforations. Such erosion can cause
inadequate zone coverage. The pad fluids in the first part of the fracturing procedure
can be effectively diverted by limited entry, but it is doubtful in many situations that
the entire treatment can be successfully diverted by limited entry zone coverage
because of the loss of perforation friction caused by the erosion of the perforations.
Perforation erosion and changing perforation friction occur by two mechanisms. The
rounding or smoothing of the entrance to a perforation causes an immediate and
significant pressure reduction. Secondly, the diameter of the perforation very slowly
increases causing a further drop in pressure versus time. Factors that affect
perforation erosion are the velocity of the fracturing fluid (pump rate), perforation size
and sand concentration.
1. Fluid Velocity. Maximum recommended flow rate through Dowell treating iron
and connections is based on a 45-ft/sec fluid velocity. This is equivalent to
8.5 bbl/min through 1.8-in.-ID treating pipe. This is considered the maximum
rate at which erosion is not excessive when pumping abrasive fluids.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 19 of 40

As a comparison, the flow rate through one 0.375-in. diameter perforation at


45 ft/sec is 0.37 bbl/min. At 1 bbl/min per perforation, not uncommon for
perforations this size, the velocity of the fluid in the perforation is 122 ft/sec and
friction pressure is 140 psi. When the proppant is added to the fluid, erosion
may be significant.
2. Perforation Size. Smaller perforations erode faster than larger ones at the same
bbl/min injection rate. To illustrate this, the flow rate of 0.37 bbl/min (from the
comparison in No. 1 above) through a 0.25-in. diameter perforation is 102 ft/sec.
At 1 bbl/min, it is 275 ft/sec and friction pressure is 700 psi.
3. Sand or bauxite slurried in water increases not only the rate of the perforation
erosion but also the amount of erosion that will occur. Higher injection rates and
higher proppant concentrations accelerate this erosion. How much erosion
takes place at a perforation is dependent on the fluid rate and proppant
concentration.
The limited entry process is effective and easy to apply. However, because of the
perforation erosion on high-rate or high-volume fracture treatments, it may be best to
use a combination of limited entry and ball sealer diversion or slurried solids to attain
diversion. These diversion techniques are discussed in later parts of this section.

2.2.2 Contributing Factors


To obtain the best selectivity and the best control of stimulation fluids with the
limited-entry technique, there should be an integrated design for the entire well
completion program. As in any conventional stimulation treatment, the amount and
type of stimulation fluids to use and the most effective injection rates can be
determined with the fracturing and acidizing computer software, and from the
laboratory analysis of formation and well-fluid samples.

2.2.2.1 Casing Size


The size and pressure rating of the casing are of prime importance in any hydraulic
stimulation treatment. This is especially true for limited entry jobs. The pipe must
not only be of sufficient size (and pressure rating) for production but also provide
sufficient size to allow the high injection rates that may be necessary for limited entry
jobs.

2.2.2.2 Number and Size of Perforations


Since access to the formation is controlled by the perforations, the size, number and
shape of the perforations are each critical variables in designing a limited entry job.
The length of the producing intervals and the number of intervals to be treated can
be selected from log or core analysis data. Using this information, the minimum
number of perforations for effective stimulation and drainage of the zones can be
determined. Assure that each zone will receive the desired amount of treatment

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 20 of 40

volume at the desired rate by proportioning the total number of perforations


according to the number and thickness of the zones.
It should be noted that limited perforations may also affect production response due
to the convergence of turbulence in the near wellbore area (pseudoskin). This
emphasizes the necessity of an integrated design for the entire well completion.
Since the shape of the perforation is a critical factor, specific perforation charges
should be selected that will maximize smooth, round and uniform perforations. The
term “perforation coefficient” is an expression of the efficiency of an orifice for
allowing fluid passage. A value of 1.00 would indicate the ideal shape while a value
of 0.95 would indicate efficient fluid passage through a smooth, burrless, tapered
inlet. A value of 0.85 is commonly used for average conditions. Fig. 11 illustrates
the change in the perforation friction versus perforation coefficient.

Fig. 11. Change in perforation friction versus perforation coefficient.


It is important to determine the number of perforations open for a limited entry
treatment. This is normally done while pumping a prepad or pad during the initial
part of the treatment. An example of this process is described in EXAMPLE 2 of this
section.
Another effective method of proportioning the fracturing fluid is to use different sizes
of perforations in different zones. This tends to complicate the perforation software,
but can contribute greatly to the ultimate success of the job.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 21 of 40

In some cases, it may not be practical to reduce the number or size of the
perforations to assure limited entry. It may be necessary to incorporate a secondary
diversion method such as ball sealers, temporary diverting agents or a combination
of these.

2.2.2.3 Differences in Fracturing Pressures


In many wells, there will be a wide variation in the bottomhole fracturing pressures of
different zones. Such variations will result in differences in the net friction pressures
across the perforations. The zone having the lowest bottomhole fracturing pressure
may accept a larger portion of the stimulation fluid. If the fracturing pressure
differences are known, it may be necessary to alter the number or size of the
perforations or the injection rate to assure the desired fluid entry into each zone.

2.2.2.4 Pressure Limitations


Normal fracturing pressure, plus the required perforation friction, plus friction in the
tubing or casing may result in surface pressures at or near the maximum allowable
for the tubulars in the wellbore.
A common belief is that the best results are obtained by maintaining the perforation
friction at a maximum during the treatment. Surface pressure normally decreases
due to
• increased hydrostatic pressure caused by adding a propping agent
• reduced friction pressure at the perforations caused by perforation erosion.
An increase in pump rate will probably be necessary to maintain the perforation
friction. This increases the likelihood of treating all perforations in the interval.

2.2.2.5 Hydraulic Horsepower Requirements


Small-diameter perforations are frequently preferred in limited entry treatments to
increase the perforation friction and, subsequently, higher pumping pressures.

2.2.2.6 Perforation Friction Pressure


Perforation friction pressures of 700 to 1000 psi are commonly used as pressure
differentials for designing limited entry jobs. Differential pressures of this magnitude
are sufficient to attain entry into most zones. However, erosion caused by a
proppant can quickly round the sharp edges of the perforations and decrease the
perforation friction. As proppant delivery is maintained, the perforation will enlarge.
This further decreases the perforation friction. A differential of 1000-psi perforation
friction may be necessary to maintain limited entry for a longer period of time. Based
on decreasing the perforation friction as proppant delivery is maintained, limited
entry is not recommended where the different zones have bottomhole fracturing
pressures of greater than 700 psi.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 22 of 40

2.2.2.7 Breakdown Technique


It is essential to the success of a limited entry treatment for all perforations to be
broken down and positive formation entry assured at all perforations. It is common
practice to perform this breakdown with acid. Three techniques have proved to be
successful and are discussed here.
1. In the ball-out method of a breakdown, ball sealers are systematically injected in
acid until a ball-out occurs.
2. Mechanical zone isolation, by using a system of packers and bridge plugs, is
another method of a breakdown. Each zone to be treated is isolated and
individually broken down. This method is time consuming but will provide
accurate bottomhole fracturing pressure data for each zone. Ball sealer
diversion, used in conjunction with this technique, is an excellent way to assure
that all perforations are open. If ball sealers are used, the bridge-plug retrieving
mandrel should be covered with sand to prevent the lodging of ball sealers
between the mandrel and the casing, thus complicating the bridge-plug retrieval.
Selective injection packers can be used to assure the perforation breakdown.
When these packers are used, a bridge plug is not required; however, ball
sealers cannot be used because of the small injection ports. Instead, injection
control valves are commonly used.
3. Matrix acidizing using a fluid-loss control agent for acid is the third method. Acid
is spotted in the wellbore prior to perforating or spotted over the perforations. It
is then displaced slowly into the formation. The primary advantage of the matrix
acidizing method is the elimination of excessive breakdown pressures which
may cause damage to the cement sheath or the opening of channels behind the
pipe. However, under the best conditions, the matrix acidizing technique will
ensure that each perforation will accept fluid at matrix rates. The technique
does not provide a breakdown of the formation at each perforation.

2.2.2.8 Determination of Formation Fracturing Pressure


This is essential to the successful design of a limited entry procedure. It may be
known from previous experience in the area, from a prior breakdown, or determined
from instantaneous shutdown pressure at the beginning of a job.

2.2.3 Design
The design of a limited entry treatment is a trial-and-error calculation. Several
different job designs may be prepared before a tentative pumping schedule meets all
of the necessary prerequisites of the surface pressure, perforation friction pressure,
pump rate, perforation size and perforation spacing. The following four examples
show how treatments may be designed for various downhole conditions. Fig. 12
pictorially represents the differences in perforating for a conventional fracturing
treatment versus perforating for a limited entry job.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 23 of 40

Fig. 12. Gamma ray-sonic log.


By using one of the four methods of calculations shown in the following example
problems, a limited entry treatment can be designed to meet almost any set of well
conditions.
At the start of any limited entry treatment, it is advisable to obtain an instantaneous
shutdown pressure. This will provide a check on the number of perforations taking
fluid. It will also permit the accurate calculation of the number of perforations taking
fluid at the time the pisi was taken. Should these calculations show that some
perforations are not taking fluid, it may become necessary to increase the injection
rate or to resort to secondary diverting methods.
Example 1
Design a limited entry treatment to stimulate eight thin porous sections within a
massive zone. Log and core analyses indicate these porous sections to each have
less than a five-foot vertical extent. The intention is to fracture stimulate each
section with a predetermined volume of fluid at a rate between 2 to 3 bbl/min per
section. The perforation diameter is to be 3/8 in. A perforation friction of between
800 and 900 psi is required.
Well Conditions  Example 1

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 24 of 40

Information From Information on Well


Offset Well to be Treated
Producing Depth 6120 to 6200 ft 6118 to 6138 ft
Casing 5-1/2 in., J55, 4-1/2 in., J55,
15.5 ppf to 6320 ft 9.5 ppf to 6282 ft
Perforations 3/8-in. diameter, 3/8-in. diameter,
4 spf, To be determined
6122 to 6200 ft
To be determined
316 holes total To be determined
Surface Treating Pressure 1350 psi Maximum Allowable =
3300 psi
Treating Fluid WF140 in 2% KCl WF140 in 2% KCl
Pump Rate 40 bbl/min To be determined

Example 1 Calculations
1. Determine the bottomhole fracturing pressure. Use the offset well information.
pw = 1350 psi
ph = (6150) (0.438) = 2694 psi
ptf = (55) (6.15) = 338 psi. (Fig. 13 provides friction pressure data for YF140
fluid.)
0.2369 q pf ρ
2
p pf =
d pf α 2
4
(1)
2
( 0.2369)(0.127 ) ( 8.43)
p pf =
( 0.375) 4 (0.85) 2
ppf = 2.2 psi
Because of the large number of perforations, injection rate per perforation is low
(0.127 bbl/perforation). Perforation friction is therefore, negligible.
piw=pw+ph-ptf-ppf (2)
piw = 1350 + 2694 - 338 - 2
piw = 3704 psi
Where:
pw = total surface treating pressure (psi)
ph = hydrostatic pressure (psi)
ptf = total tubing or casing friction pressure (psi)

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 25 of 40

ppf = perforation friction pressure (psi)


qpf = injection rate per perforation (bbl/min)
p = fracturing fluid density (lbm/gal)
dpf = perforation diameter (in.)
α = perforation coefficient (dimensionless, 0.85 is commonly used)
piw = bottomhole fracturing pressure (psi).

Fig. 13. Friction of WF140 in 4 1/2 and 5 1/2 in. pipe.


2. Determine Pump Rate per Perforation at a Perforation Friction Pressure of 800 to
900 psi  use Eq. 1.
Trial and Error: How much perforation friction will be generated at (a) 2, (b) 2.5,
and (c) 3 bbl/min per 3/8-in. diameter perforation? Each of these rates is in the
desired range for treating the eight perforated sections.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 26 of 40

a. How much perforation friction will be generated at 2 bbl/min per perforation?


Use Eq. 1.

0.2369 q pf ρ (0.2369)( 2)2 ( 8.43)


2
p pf = =
d pf α 2
4
(0.375) 4 ( 0.85) 2

ppf = 559 psi


This is an insufficient perforation friction pressure.
b. How much perforation friction will be generated at 2.5 bbl/min per perforation?
( 0.2369)( 2.5) 2 ( 8.43)
p pf =
(0.375) 4 ( 0.85) 2
ppf = 874 psi
This is an acceptable perforation friction pressure.
How much perforation friction will be generated at 3 bbl/min per perforation?
( 0.2369 )( 3) 2 ( 8.43)
p pf =
(0.375) 4 ( 0.85) 2
ppf = 1258 psi
This is more perforation friction pressure than requested.
Conclusion  A perforation friction pressure of 874 psi (2.5 bbl/min per
perforation) is acceptable for this job.
Assumption  If one 3/8-in. diameter perforation is shot in the center of each
of the eight intervals and the required pump rate is 2.5 bbl/min per perforation,
the total pump rate will be 20 bbl/min.
Can the well be fracture treated at 20 bbl/min and stay under the 4000-psi
maximum allowable surface pressure?
3. Determine the Total Surface Pumping Pressure for the Well to be Treated, using
ppf = 874 psi and 20-bbl/min total pumping rate. Rearrange Eq. 1.
pw = piw + ptf + ppf - ph
piw= 3687 psi from the offset well.
6151
ptf = ( 55) = 338 psi
1000
ppf= 874 psi
ph= (6151)(0.438) = 2694 psi
pw = 3687 + 338 + 874 - 2694
pw = 2205 psi

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 27 of 40

4. Determine Hydraulic Horsepower Requirements  Eq. 3.


pw q
HHP =
40.8
(3)
( 2205)( 20)
HHP =
( 40.8 )
HHP = 1081

This well can be treated using the limited entry technique satisfying all pressure and
rate restrictions at 20 bbl/min through eight 3/8-in. perforations. Also, hydraulic
horsepower requirements are minimal because the casing friction pressure is
minimal.
Example 2
Calculate the number of perforations needed for limited entry, the pumping rate
required and the surface pressure expected to fracture stimulate a formation having
a long vertical extent. Use the same well conditions as in Example 1 for the well to
be treated but with the following exceptions:
• Log information and core analysis indicate uniform porosity and permeability
through the entire section between 6118 to 6138 ft.
• Maximum allowable surface treating pressure is 3300 psi.
Well Conditions  Example 2

Information From Information on Well


Offset Well to be Treated
Producing Depth 6120 to 6200 ft 6118 to 6138 ft
Casing 5-1/2 in., J55, 4-1/2 in., J55,
15.5 ppf to 6320 ft 9.5 ppf to 6282 ft
Perforations 3/8-in. diameter, 3/8-in. diameter,
4 spf, To be determined
6122 to 6200 ft To be determined
Surface Treating Pressure 1350 psi Maximum Allowable = 3300 psi
Treating Fluid WF140 in 2% KCl WF140 in 2% KCl
Pump Rate 40 bbl/min To be determined

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 28 of 40

Example 2 Calculations
1. Determine the allowable total friction pressure available to use in this well using
the 3300-psi total treating pressure as a base. First, solve for ptf + ppf.
ptf + ppf = pw + ph - piw
= 3300 + 2694 - 3673
= 2321 psi
Total pipe and perforation friction available for use is 2321 psi.
2. Assume a reasonable perforation friction. A perforation friction of 900 psi is used
here. It follows, using Eq. 1, that the friction pressure available as pipe friction is
1421 psi.
ptf = 2321 - 900
= 1421 psi
ptf 1421
= × 1000 = 231 psi / 1000 ft
D 6151
3. A rate of 58 bbl/min can be attained with a 230-psi friction loss per 1000 ft in the
casing (Fig. 11).
4. Determine the flow rate through a 3/8-in. diameter perforation that will attain
approximately 900 psi. Solve Eq. 1 for qpf.
 p pf d pf 4α 2 
q pf = 1/ 2
 0.2369ρ 
 
 ( 900)(0.375) 4 ( 0.85 ) 2 
q pf =   1 / 2
 ( 0.2369 )(8.43) 2 
= 2.5 bbl/min per perforation
5. Calculate the total number of perforations that could be used under these
conditions.
58 bbl / min
Number of Perforations = = 23
2.5 bbl / min / Perf
6. Thus, an effective limited entry fracture stimulation may be obtained with as many
as twenty-three 3/8-in. diameter perforations at 58 bbl/min and 3300-psi surface
pressure. Hydraulic horsepower requirements (Eq. 3) would be:

pw q ( 3300 )( 58 )
HHP = = = 4691
40.8 40.8

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 29 of 40

Example 3
Problem  Calculate the number of perforations open and accepting fracturing fluid
during the pad portion of the fracturing treatment. Use the same well conditions as
in Example 2, but with the following.
• Assume that the fracture treatment has started and, while the pad fluid was being
injected out the perforations on the bottom and prior to adding sand at the
surface, an instantaneous shutdown pressure (ISIP) was taken. Pump rate just
prior to shut down was 45 bbl/min. Surface treating pressure was 3300 psi. The
ISIP was 1050 psi.
Well Conditions  Example 3

Information From Offset Information on Well


Well to be Treated
Casing 5-1/2 in., 13.5 ppf, J55 set 4-1/2 in., 9.5 ppf J55 set to
to 6202 ft 6198 ft
Fracturing Fluid WF140 in 9 lbm/gal salt WF140 in 9 lbm/gal salt water
water
Producing Zone A
Depth 6110 to 6156 ft 6100 to 6120 ft
Perforations 0.375 in., 4 spf, To be determined
6112 to 6149 ft
Treating Pressure 1350 psi To be determined
Injection Rate 30 bbl/min To be determined
Producing Zone B
Depth 5890 to 5926 ft 5880 to 5900 ft
Perforations 0.375 in., 4 spf, 5894 to To be determined
5922 ft
Treating Pressure 1200 psi Maximum Allowable = 3500 psi
Injection Rate 40 bbl/min To be determined

Example 3 Calculations
1. Determine the total friction being used at 45 bbl/min. Rearrange and use Eq. 2.
ptf+ ppf = pw + ph - piw
= 3300 + 2694 - 3673
= 2321 psi

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 30 of 40

2. At 45 bbl/min, the pipe friction is 150 psi/1000 ft. Calculate the total pipe friction.
(150)(6151)
ptf = = 923 psi
1000
3. Calculate the total perforation friction.
ppf = 2321 - 923 = 1398 psi
4. Calculate the pump rate per perforation. Use Eq. 1 as in Example 2, Item 4).
 (1398)(0.375) 4 ( 0.85) 2 
q pf =   1 / 2
 ( 0.2369 )(8.43) 
= 3.16 bbl/min
Pump rate per perforation is 3.16 bbl/min.
5. Determine the number of perforations open and accepting fluid.
Total Pump Rate
= Number of Perforations Open
Pump Rate Per Perforation
45 bbl / min
= 14.2 = 14 or 15 perforations are open.
3.16 bbl / min / Perf
6. There may be several possible reasons for the apparent low number of open
perforations.
a. The bottomhole fracturing pressure is slightly higher than the offset well,
3673-psi for the offset and 3823 psi for the well being treated. This might
have some effect.
b. This well was not broken down and balled off prior to the fracture treatment.
A ball-off job prior to a limited entry fracturing treatment is a must.
c. There may be metal burrs and sharp angles at or in the perforations adding to
the perforation friction. Any additional friction from this source should quickly
decrease when the proppant reaches the perforations.
Example 4
Determine the perforation pattern required to fracture stimulate, using limited entry,
two different zones that have different bottomhole fracturing pressures. The zones
are to be simultaneously treated.
Offset: Each zone was treated separately, using a retrievable bridge plug.
Well to be Treated: Log information and core analyses indicate that it is desirable to
fracture stimulate Zone A from 6100 to 6110 ft and Zone B from 5880 to 5890 ft
(10 ft in each zone). Plan for a pump rate for the fracturing treatment to be in the
10-bbl/min range.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 31 of 40

Example 4 Calculations
1. Determine Zone A and Zone B fracturing pressures from the offset well.
Zone A
piw =pw+ph-ptf-ppf (2)
pw = 1350 psi
ph = (0.468) (6130) = 2869 psi
ptf = 30 psi/1000 ft = (30) (6.130) = 184 psi
ppf = negligible due to pump rate per perforation.
30 bbl/min/(4 spf x 38 ft) = 0.2 bbl/min/perforation. Eq. 2 can be used to check
this.
piw = 1350 + 2869 - 184
= 4035 psi
Zone B
piw = 1200 + (0.468)(5908) - (55)(5.908)
= 3640 psi
Zone A has a bottomhole fracturing pressure of 395 psi (400 psi) higher than Zone
B. Thus, the perforation friction pressure drop on Zone B must be 395 psi higher
than on Zone A. If a desired perforation friction of 900 psi is maintained on Zone A
during the treatment, the effective perforation friction at Zone B will be 1300 psi.
2. The problem now becomes one of trial and error; how many of what size
perforations must be shot in each zone to provide a 1300-psi friction pressure
across Zone A and a 900-psi friction pressure across Zone B as well as provide
the same injection rate into each zone? The injection rate for each zone is to be
near 10 bbl/min. Check with Schlumberger Wireline to determine the most
common perforation sizes used locally. This example uses 0.25, 0.325, 0.375,
and 0.50-in. diameter perforations.
Flow Rate per 0.25-in. Perforation
At 900 psi = 1.09 bbl/min (using Eq. 1)
At 1300 psi = 1.31 bbl/min
Flow Rate per 0.325-in. Perforation
At 900 psi = 1.84 bbl/min
At 1300 psi = 2.22 bbl/min
Flow Rate per 0.375-in. Perforation
At 900 psi = 2.46 bbl/min
At 1300 psi = 2.95 bbl/min

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 32 of 40

Flow Rate per 0.50-in. Perforation


At 900 psi = 4.37 bbl/min
At 1300 psi = 5.25 bbl/min
3. Using the above values, construct a table showing the perforation diameter,
pressure, rate and number of perforations required for approximately 10 bbl/min
per zone.
4. Make the best match of rates at 900 psi and 1300 psi. The best match in this
example is 8.74 bbl/min through two 0.50-in. perforations and 8.85 bbl/min
through three 0.375-in. perforations.
Therefore, three 0.375-in. perforations should be shot in Zone A and two 0.5-in.
perforations should be shot in Zone B. Total pump rate for both zones should be
18 bbl/min.
5. Calculate the pertinent treatment parameters.
pw = piw + ptf + ppf - ph
D = 5995 ft
ppf = 1300 psi (maximum)
piw = 4035 psi (maximum)
ph = (9)(0.052)(5995) = 2806 psi
ptf = (45)(5.995) = 270 psi
pw = 4035 + 270 + 1300 - 2806 = 2799 = 2800 psi
pw q ( 2800 )(18)
HHP = = = 1235 HHP
40.8 40.8

Table 2. Example for Selecting a Perforation Schedule for Multiple Zones


Having Different Bottomhole Fracturing Pressures
Perforation Injection Rate/Perforation Number of Perforations at
Diameter (in.) Approximately 10 bbl/min
900 psi 1300 psi 900 psi 1300 psi
(bbl/min) (bbl/min) (bbl/min) (bbl/min)
0.25 1.09 1.31 10 to 10.90 8 to 10.50
0.325 1.84 2.22 5 to 9.20 4 to 8.88
0.375 2.46 2.95 4 to 9.84 3 to 8.85
0.50 4.37 5.25 2 to 8.74 2 to 10.50

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 33 of 40

2.3 Single-Point Entry


The single-point entry (injection into one zone at a time) technique is a modification
of limited entry and normal ball sealer diversion. It affirms positive diversion and
assures that only one zone at a time is treated.
The technique requires perforating each zone with an equal number of perforations,
with the total surface area of perforations in each zone being sufficient to accept all
of the designed injection rates with no more than approximately 100-psi differential
pressure across the perforations. In this way, treatment of only one zone at a time is
assured. There must be a small enough number of perforations in each zone that
they can be effectively sealed with an exact number of ball sealers, achieving
positive diversion.
It is necessary to accurately know the size and shape of the perforations to make the
limited entry and any modified limited entry treatments, that is, “single-point entry”
work. Thus, specific perforating guns must be used which provide uniform
perforations.
Having properly perforated the various zones in the well, they are to be broken down
(pumped into at a fracturing rate) to ensure that the perforations are all open. If
some of the perforations will not accept fluid, all zones will not be treated as
designed and the objective of the fracture treatment will not be fully accomplished.
As in limited entry, a ball-out job using acid as the carrier fluid for the ball sealers is
the best assurance that all perforations are open. The volume of acid frequently
used is about one barrel per perforation.
After breakdown, to separate the fracture stages, the exact number of ball sealers is
injected that will seal all of the perforations in one zone. Repeat the process for
each zone until all have been fractured.
This technique works best where each zone exhibits a different fracturing pressure
and each zone is of approximately equal thickness. It should be recognized that
different volumes can be placed in different zones only when the fracturing pressure
of each zone is known.
Single-point entry may also affect production response due to the convergence of
turbulence in the near well-bore area (pseudoskin). This emphasizes the necessity
of an integrated design for the entire well completion.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 34 of 40

2.4 Plugback (Pine Island) Technique


Another treatment diversion technique, developed by AMOCO in the Pine Island field
of Louisiana, uses sand or gravel to plug back above a zone previously fractured
prior to perforating and fracturing the next zone up the wellbore. Fig. 14 illustrates
how this procedure is applied. In effect, the plugback technique is a series of
separate fracture treatments.

Fig. 14. Plugback Pine Island technique.


A typical well may have three zones to be fracture stimulated. The lower interval can
be perforated and fracture treated. Then a sand plug is placed over the first zone.
After the sand plug has been pressure tested, the second interval can be perforated
and fracture stimulated. The second zone would also be covered with a sand plug,
and the third zone would be perforated and stimulated. After the third fracture
treatment, the wellbore can be cleaned out with coiled tubing or a conventional
workover rig.
The sand plug itself is normally made of 20/40-mesh sand. In sand fill-up
calculations, the bulk density of the sand (14.3 lbm/gal or 1714.3 kg/m3) must be
used. The following guidelines can be used for selecting the amount of sand for
plugback.
• lbm sand per linear ft = (14.3)(gal/linear ft)
• kg/m = (1714.3 kg/m3)(m3/m)

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 35 of 40

also
• Linear ft/lbm sand = (0.07)(linear ft/gal)
• m/kg = (0.00058 m3/kg)(m/m3)
also
• Additional information and tables about sand fill-up in the openhole, casing,
tubing and annuli is provided in the Dowell Field Data Handbook.
Sand can be made more efficient by using a mixture 20/40 and 100-mesh sand to
reduce the permeability of the sand in the wellbore. Some fluid may initially leak
through the plug, but the addition of fluid-loss additives to the following fluids should
quickly seal the plug and prevent fluid from migrating through the sand.
Many different versions of the plugback technique can be applied. For instance,
each zone could be fractured and cleaned up before setting the next sand plug. If all
of the different stages can be pumped in one or two days, that is, if the fractured
zones can be shut in for this length of time without being damaged because of the
presence of large amounts of treatment fluids, all zones could conceivably be
cleaned up simultaneously after all treatment stages have been pumped.
The plugback placement can be enhanced by circulating the sand into position rather
than allowing extra time for the sand plug to gravitate through viscous fracturing
fluids and possibly collect on pipe collars.
As with the other treatment diversion techniques, problems may develop. The
following lists some problems that can be encountered.
• It is sometimes very difficult to place the sand plug precisely where it is needed.
If too much sand is left in the wellbore, it could cover up part of the next zone to
be perforated and stimulated; therefore, the excess sand would have to be
washed from the wellbore.
• In other cases, the sand plug tends to move downhole or through the perforations
to void places behind the pipe. Thus, extra time is required to dump additional
sand and allow it to settle at the proper position.
• Movement of the isolating sand plug while attempting the flowback of the uphole
interval can cause damage to the surface equipment.
• Coiled tubing, used in cleanup operations, can stick if there is excessive sand-
plug movement.
• Damage to the production can be caused by fluid invasion from cleanout
operations or from being shut in too long after the fracture treatment.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 36 of 40

2.5 Slurried Solids


These are solid particles (flake, granular, powder) which are slurried in fluid prior to
injection into the well during a fracture treatment. They function in a similar manner
to a drilling mud which contains lost circulation material. The basic difference is that
the lost circulation material in drilling mud can act slowly to obtain shutoff, but
slurried solids diverters must act instantly in fracture stimulation treatments.
In continuous staged fracturing treatments, each stage is identical and each zone
receives its own pad volume and graduated proppant schedule. At the end of each
stage, enough diverter is injected to either block the perforations of one zone or
bridge in the fracture of the zone being treated. The plan is to effectively divert the
next stage into another set of perforations and open a new fracture opposite these
perforations. The diverter is then followed by the pad volume to initiate the fracture
in the next zone and the process is repeated. The diverter is eliminated at the
completion of the final proppant stage.
Dowell markets different formulations and mixtures of diverting agents under the
trade name of FIXAFRAC* Services.
Below is a partial list of slurried solids diverting agents, shows that a diverting agent
for fracture treatments is typically a graded material that is insoluble in fracturing
fluids but soluble in formation fluids.
Table 3. Diverting Agents
• Coarse Rock Salt
• Graded Rock Salt
• Graded Paraformaldehyde
• Flake Benzoic Acid (Fine And Coarse)
• Graded Oil-Soluble Resin (Low Temperature)
• Graded Oil-Soluble Resin (High Temperature)
• Wax Beads
• Mixture of Oil-Soluble Resin and Gum Karaya
• Graded Naphthalene
• Oyster Shells
• Polymer-Coated Sand (see BRACKETFRAC* Service, Section 7.5)
• Buoyant Particles (see BRACKETFRAC Service, Section 7.5)
• Flake Boric Acid

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 37 of 40

The following define appropriate conditions for a slurried solid to behave as a


diverting agent.
• The largest particle size of the diverter must exceed the width of any fracture to
be bridged. This will reduce the rate of fluid being injected into the fracture. For
successful diversion, these large particles must be followed by successively
smaller particles so that the fracturing fluid will be diverted and will fracture
another zone.
• The amount of differential pressure held by any material depends on the strength
of the largest particle.
• Some particle deformation of the diverting material is desirable, so long as it does
not lead to structural failure of the particle.
• A fluid-loss-control additive and a material to cause the particles to adhere to
each other will greatly enhance the efficiency of the diverter. Without the smaller
particles, complete shutoff is not possible. There is an optimum blend of larger
and smaller particles.
• A viscous fluid is needed to carry the diverter solids, but too much viscosity is
detrimental to the bridging and plugging process.
• The most effective agents are those that are insoluble in their carrying fluid.
However, they should either be soluble in the fluids produced from the well or
else be unstable at bottomhole conditions so that after a few days of exposure,
they will be removed during production.
• The amount of diverting agent needed for diversion depends on the width and
height of the fracture at the wellbore, differential pressure applied, particle-size
distribution of the solids in the diverter and fluid-flow velocity past the agent. The
amount of diverter to use in a given well is determined by experience and by trial
and error. If multiple stages of slurried diverters are to be run, the amount to use
on the successive stages depends on the effectiveness of the first stage of the
diverter, that is, “should more, less or the same amount of diverter be run to
obtain appropriate surface pressure indication of diversion?”
The slurried solids technique does not perform efficiently for diversion when high-
viscosity fracturing fluids are used to carry the propping agents. High-viscosity
fracturing fluids (crosslinked fluids) normally create very wide fractures and the
slurried solids will not successfully seal these fractures near the wellbore. Therefore,
slurried solids are expected to be only minimally successful in diverting such fracture
treatments.

2.6 Viscous Fluids


Viscous fluid diverters and viscous pill diverters can be placed during continuously
pumping fracture operations but require specific shutdown times to allow the
viscosity diverter to congeal in the mouth of the fracture and gel into a rigid mass.
Gelation is time, temperature and chemical dependent and can be adjusted to fit the

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 38 of 40

treatment and well conditions. After the gel has become rigid, the next stage of the
fracture treatment can be performed. Internal chemical breakers are used to cause
the gelled diverters to break down and return to a thin fluid state in a predetermined
time at the bottomhole static temperature of the well.
Careful engineering and strict quality control are mandatory.
PROTECTOZONE* Services have been used successfully for this purpose. For
information regarding gelation times, breakdown times, fluid loss, formation
damage, volumes of gel needed and mixing, refer to the Matrix Materials Manual.
A temperature survey, radioactive survey or other appropriate electric log is
frequently run after the staged fracture treatment for assurance that fracture
diversion was attained.

2.7 Baffles and Balls


The casing baffle/fracturing ball diversion technique has been widely used in shallow
gas wells of Appalachia and has also been applied in other shallow gas areas.
Fig. 15 illustrates a two-stage fracturing procedure that uses this baffle and ball
technique.
• The baffle is run with the casing and is placed between two zones that are to be
fracture treated.
• The lower zone is perforated and fracture stimulated.
• A fracturing ball is dropped to seal in the baffle.
• The upper zone is perforated and fracture stimulated.
• Flowing the well will remove the fracturing ball, and the zones can be
simultaneously produced.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Diverting Techniques
Page 39 of 40

Fig. 15. Two-stage fracturing procedure, using baffle and ball technique.
Although Fig. 15 illustrates the casing baffle/fracturing ball diversion between two
zones, this technique can be used to selectively treat more than two zones (by
running more baffles and dropping more balls). The baffles are placed between the
three or more zones to be fracture treated, and assurance is made that the holes in
the casing baffles are progressively smaller with increasing depth. Fracture
diversion is attained by dropping, down the casing (between the fracturing stages),
progressively larger fracturing balls to correspondingly seal the larger baffles.
The advantage of using this diversion technique is positive diversion without rig time.
However, in instances where the fracturing ball will not flow back to the surface, a rig
is necessary to drill up the ball and baffle. A normal rock bit is commonly used.
Where these situations are anticipated, it is best to use a “bakelite type” fracturing
ball so that it will be more likely to shatter when contacted by the drill bit. Nylon
fracturing balls, tougher than the bakelite type, are available.
All casing baffle plates have an inherent weakness due to the limited area available
in the casing coupling. Pressure differential across the casing baffle should be
limited to the manufacturer's specification. Use of casing baffles in pipe sizes
greater than 5 1/2 in. is not recommended because the total force applied to the
cross-sectional area of the baffle may exceed the strength of the baffle.

DOWELL CONFIDENTIAL
Section 700.7
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Diverting Techniques Dowell
Page 40 of 40

2.8 Bridge Plugs


The most reliable technique for diverting fracture treatments into multiple intervals is
the bridge-plug method (also called the packer-and-bridge-plug method). A typical
treatment using this method involves perforating a lower zone, performing a fracture
treatment on that zone, setting a bridge plug above that particular interval, and then
perforating and treating the next zone up the wellbore. The process can be repeated
several times before the well is cleaned out and all of the zones previously fracture
stimulated are commingled.
Some advantages of the bridge-plug method are:
• it is the most positive way to divert a fracture treatment
• perforations can be tailored to each specific zone
• the size and rate of each fracture-treatment stage can be tailored to each specific
zone (when performed correctly, each zone should be optimally stimulated and
income from the production should be maximized).
The disadvantages of the bridge plug method include
• a workover rig is required to make multiple moves of the packer and bridge plugs
in the well
• the method is time consuming
• mechanical problems can occur with the retrievable bridge plugs used to isolate
the intervals
• formation damage (water block, water-sensitive clays or other saturation change)
can be caused by long shut-in times
• the method can be costly, particularly when a workover rig is required.
Coiled tubing is frequently used to run, set and retrieve bridge plugs and also for well
cleanout after fracturing. This normally represents a cost savings over the necessity
of having a workover rig on the well.
If the pressure differential across a retrievable bridge plug becomes too large, then a
drillable plug must be set between the zones to be fracture treated. Under most
conditions, the drillable plugs can be easily removed; however, situations could
develop where the drillable plugs become difficult to remove from the well. It is very
undesirable to drill a plug and circulate a well with an open fracture above the plug.
Lost circulation particles or drilling fines can damage the exposed fracture(s).
The cost of fracture diversion should not be the sole indicator of the most economical
diversion technique. Use of the FracNPV software as a function of the treatment
diversion technique should be studied. It is quite likely that the most expensive
diverting technique is also the most economical diverting technique, because it will
improve the cumulative production and cash flow.

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFLOW Technology
Page 1 of 7

CleanFLOW TECHNOLOGY

1 Introductory Summary............................................................................................................. 1

2 The CleanFLOW System ......................................................................................................... 2


2.1 Optimized Breaker Composition .......................................................................................... 2
2.2 J495 CleanFLOW Additive................................................................................................... 3
2.3 CleanFLOW Application....................................................................................................... 4
2.3.1 Temperature ............................................................................................................... 4
2.3.2 J495 Concentrations................................................................................................... 4
2.3.3 Breaker Interactions ................................................................................................... 4
2.3.4 Material Handling........................................................................................................ 5
2.3.5 Field Results ............................................................................................................... 5

3 Conductivity Measurements ................................................................................................... 5


3.1 Polymer Concentration ........................................................................................................ 6
3.2 Pressure-Drop Measurements............................................................................................. 7
3.3 Other Parameters ................................................................................................................ 7
3.4 Summary.............................................................................................................................. 7
TABLES
Table 1. Breaker Schedule Design Example ............................................................................... 3

1 Introductory Summary
The CleanFLOW* technology is a result of a research focus on increased
understanding of proppant-pack damage and cleanup. A recent study by Dowell
research has correlated theoretical flow through porous media (using the Kozeny-
Carmen equation) with proppant-pack permeability measurement (using Darcy’s
law). The result is a relationship between reduction in porosity and retained
permeability.
The most dramatic result from this correlation is the demonstration of an exponential
drop in retained permeability for only a small reduction in porosity (assumptions
include low Reynolds number and randomly distributed pore blockage). A 10%

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFLOW Technology Dowell
Page 2 of 7

reduction in porosity (for example, 30% down to 27%) results in a 35% reduction in
retained permeability (for example, 100% down to 65%). Retained permeability
values of 85 to 100% require the residual porosity reduction to be less than 5%.
Additionally, the fractional porosity lost at the core face due to polymer concentration
and filtercake formation results in a significant retained-permeability loss. Laboratory
simulation supports this correlation.
A major conclusion from this research is that small, random pore blockages by
residual polymer or even a thin, concentrated polymer filtercake at the fracture face
can result in large retained-permeability decreases. Effective means of
removing/destroying the polymer residue are needed, both in the propped fracture
and at the wall.
A more complete treatise of this work is provided in Society of Petroleum Engineers
(SPE) 28511, which also deals with the effects of viscous fingering through the
proppant pack. Combined with the additional factors described in the paper (mobility
ratio, polymer concentration, weight average molecular weight [Mw]), there is an
inherent advantage to breaking down the polymer. However, further experimental
investigation showed that decreases in Mw below 100,000 to 200,000 did not
necessarily improve proppant-pack cleanup. Polymer with a lower Mw resulted in
blocked porosity not explained by the mobility ratio concept. The CleanFLOW
technology was developed to address this degraded polymer damage.

2 The CleanFLOW System

2.1 Optimized Breaker Composition


The foundation of the CleanFLOW system is a properly optimized breaker; a breaker
designed to function at the desired temperature range for a period of time long
enough to facilitate adequate polymer degradation. A comparison of the most recent
addition to the EB-Clean* breaker family (J496 Encapsulated Breaker) with the
previous generation persulfate breakers shows a significant enhancement in thermal
stability (half-life) at temperatures greater than 180°F (82°C).
Activity profiles of the various oxidizing breakers at temperatures ranging from 125 to
275°F (52 to 135°C) show that each breaker has an optimum region of performance,
where activity, control and cost are balanced. J496 is designed for use at
temperatures ranging from 180 to 235°F (82 to 113°C). It can be used in most fluids
in this temperature range, with the exception of the new borate-crosslinked fluid,
YF100.1HTD. In this temperature range combinations of Breaker J218, EB-Clean
J475 Breaker and Breaker J481 have also been used.
The design of the breaker to provide optimum activity in the desired temperature
range means more breaker in the proppant pack for a longer period of time,
enhancing cleanup and improving retained permeability. The optimization of J496 is

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFLOW Technology
Page 3 of 7

based on changes in both the core chemistry (to a non-persulfate core) and in the
optimization of the encapsulation to allow for optimum delay at these temperatures.
Because of decreased activity at lower temperatures and the effects of near-wellbore
cooldown, an optimum design at temperatures ranging from 180 to 200°F (82 to
93°C) can include J496 in the pad and early proppant stages, followed by J475 with
or without J218 in the later proppant stages. Table 1 provides an example.

Table 1. Breaker Schedule Design Example


Stage Fluid J218 Conc. J475 Conc. J496 Conc. J495 Conc.
Volume(gal) (lbm/1000 gal) (lbm/1000 gal) (lbm/1000 gal) (gal/1000
gal)
Prepad 20,000 0 0 0 2
Pad 15,000 0 0 6.7 2
3 PPA 7,500 0 1 4 2
4 PPA 7,500 0 1 4 2
6 PPA 10,000 0 1 4 2
8 PPA 10,000 0 3 0 2
10 PPA 5,000 0 3 0 2
Flush 7,097 2 0 0 0
Note: J495 CleanFLOW Additive improves breaker performance.

2.2 J495 CleanFLOW Additive


The second part of the CleanFLOW system is the proprietary additive blend J495.
J495 aids the breaker in proppant-pack cleanup by interacting with broken polymer
fragments. Repeated conductivity testing has shown that degradation of the polymer
backbone alone will not necessarily cause improved conductivity. In many cases an
increase in breaker concentration actually causes a decrease in conductivity. As the
temperature increases above 200°F (93°C), the degraded polymer effect becomes
more severe. This effect has been attributed to a chemical modification of the
degraded polymer fragments at elevated temperatures, which causes an increase in
their tendency towards pore plugging.
The dispersive action of the J495 enhances the flow of the broken polymer out of the
proppant pack, thereby increasing available flowpaths and resultant permeability.
The improvements in conductivity vary from incremental to dramatic, depending on
the temperature of application.
Accurate terms for describing J495 are ‘anticoagulant’ or ‘dispersant’, since these
are the effects of the additive on degraded polymer removal. Conversely, J495 does
not significantly alter other fluid characteristics (it is essentially a neutral water-

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFLOW Technology Dowell
Page 4 of 7

wetting agent), and it should not be used as a replacement for any of the fracturing
fluid additives specifically designed for surface-tension modification or foaming.

2.3 CleanFLOW Application

2.3.1 Temperature
J495 can be used with the appropriate breaker or breakers at temperatures ranging
from 125 to 275°F (52 to 135°C). Below this temperature range, little if any
improvement has been seen with the additive, and the interaction with Liquid
Breaker Aid J318 has not been quantified. Above this temperature range, J495
experiences gradual thermal degradation, thereby limiting its effectiveness.
Temperatures Ranging From 125 to 180°°F (52 to 82°°C)
At temperatures ranging from 125 to 180°F (52 to 82°C), the CleanFLOW additive
can be used in conjunction with J218 and/or J475. The conductivity improvement
can be as much as 50% greater than breaker-only systems. The improvement is
concentration dependent, with J495 concentrations more than 4 gal/1000 gal
providing only marginal improvements in conductivity that are not cost effective.
Temperatures Ranging From 180 to 235°°F (82 to 113°°C)
At temperatures ranging from 180 to 235°F (82 to 113°C), the CleanFLOW additive
can be used with J496 in all fluids except for YF100.1HTD, or with J218/J475 (lower
end of the temperature range) or J481/EB-Clean J490 HT Encapsulated Breaker
(high end of the temperature range). The conductivity improvement is almost 50%
greater than the previous generation of breaker-only systems (J218/J475), and an
improvement of 80 to 90% can be obtained with J495 concentrations as high as
4 gal/1000 gal.
Temperatures Ranging From 235 to 275°°F (113 to 135°°C)
At temperatures ranging from 235 to 275°F (113 to 135°C), the CleanFLOW additive
can be used in conjunction with J481 and/or J490. The damage difference
classically seen between zirconium and borate crosslinkers is eliminated with the
CleanFLOW additive in this temperature range.

2.3.2 J495 Concentrations


The CleanFLOW system effectiveness can be enhanced by increasing the additive
concentration, but eventually the application becomes less cost effective. J495
concentrations greater than 4 gal/1000 gal are probably not cost effective at any
temperature. Concentrations of 2 gal/1000 gal are generally recommended at
temperatures less than 180°F (82°C).

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFLOW Technology
Page 5 of 7

2.3.3 Breaker Interactions


J495 has a slight ‘scavenging’ interaction with J218 and J475. The J495
CleanFLOW ADDITIVE manual section in the FRACTURING MATERIALS MANUAL
- ADDITIVES provides additional information.
J495 is not compatible with Breaker J134 or Enzyme Breaker J134L. It also is not
compatible with the crosslink systems used in the YF300LPH and YF400LPH fluids.

2.3.4 Material Handling


The J496 ENCAPSULATED BREAKER manual section in the FRACTURING
MATERIALS MANUAL - ADDITIVES provides material handling/addition procedures.
J495 does not require any special handling. Typically, J495 is added using a liquid-
additive pump on the POD* blender.

2.3.5 Field Results


The CleanFLOW system has been used in numerous treatments at temperatures
ranging from 180 to 270°F (82 to 132°C). Polymer flowback comparisons have been
used as a quantitative measure of proppant-pack cleanup. If a well using the
CleanFLOW system returns a higher level of polymer than an offset well using other
breaker technology, it should indicate better proppant-pack cleanup, potentially
resulting in higher production.
Field results have shown higher polymer returns in wells using the CleanFLOW
system, in comparison to offset wells using breaker-only technology. In western
Wyoming, a well where the CleanFLOW technology was applied returned
approximately 60 to 65% of the polymer in the volume of fluid returned, with
indication that more polymer was continuing to be returned following the end of
sampling based on the gradual slope of the curve. This compares to only 30 to 35%
of the polymer being returned in a breaker-only well, with the slope of the return
curve indicating little if any additional polymer was returned after the sampling period
ended. In many cases, this higher polymer return has resulted in higher production
as well.

3 Conductivity Measurements
Much work has been performed to understand the large variability in conductivity
data being published for seemingly similar fluid systems. The attempt here is not to
discredit one set of data over another, but to explain the critical differences in testing
that can help account for large discrepancies in results.
All Conductivity Test Methods are not Alike
All conductivity tests are not alike. It is vitally important to understand the major
factors that affect the outcome of a conductivity test so that an informed decision can

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
CleanFLOW Technology Dowell
Page 6 of 7

be made when trying to compare different fluid system results. American Petroleum
Institute (API) RP 61 is the most recent API specification for conductivity testing, but
it falls far short of specifying the parameters necessary to ensure reproducible fluid
test results. Dowell has invested years of development resources to devising and
optimizing conductivity testing. The major contract laboratories that routinely
perform conductivity testing, and for which some interlaboratory comparison testing
has been done, include StimLab, Inc. in Duncan, OK and FracTech, Ltd. in the UK.
A typical conductivity test, using modified API cells and the testing procedures
detailed in SPE 16900 (StimLab procedure), involves dozens of steps, each having
at least a dozen variations, depending on the parameters being studied. (Copies of
the Dowell-recommended procedure are available for distribution.) Some of the
more prominent and widely discussed parameters are:
• closure stress - timeframes
• flowback fluid - ionic factors, dissolved gases
• flowrate control
• pressure drop measurements - procedures, equipment
• polymer concentration
• core flow versus pack flow
• method of filter cake deposition.
In an attempt to systematically determine which of these parameters were most
critical to obtaining reproducible conductivity test results, Dowell initiated a year-long
study that included interlaboratory comparisons. This involved multiple tests with
numerous test parameter variations at Dowell (including cell design), followed by
comparison tests at both major conductivity contract laboratories (StimLab and
FracTech).
The results showed that polymer concentration, the method of pressure drop
measurements and flowrate regulation (especially during the first 10 minutes), were
extremely critical to obtaining consistent results. The method of filter cake deposition
could not be distinguished from the random testing variation. Flowback fluid
preparation and direction of pack flow had a synergistic effect on results, but were
indistinguishable when considered separately in a limited number of tests. The
understanding of these results is a potential focus of future conductivity testing
research.

3.1 Polymer Concentration


Polymer concentration effects on conductivity are nonlinear, with a greater variation
(±50% or more) occurring at polymer concentrations less than 200 lbm/1000 gal.
Variation in the polymer concentration factor is viewed as being one of the significant
reasons for past discrepancies in reported conductivity numbers. A study of 365
North American fracturing treatments showed that approximately 70% had polymer

DOWELL CONFIDENTIAL
Section 700.8
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
CleanFLOW Technology
Page 7 of 7

concentration factors of at least 10. For this reason, a 10-fold concentration factor is
used as a baseline condition for Dowell conductivity tests. The method of fluid
loading and static filtercake deposition, to give a quantitative gel/proppant ratio and
subsequent polymer concentration, are parts of the Dowell-recommended
conductivity test procedure.

3.2 Pressure-Drop Measurements


Typical pressure-drop measurements during a conductivity test are extremely small,
going down to less than 0.01 psi for a minimally damaged pack. Accuracy and
linearity of these measurements must be insured through consistent quality control
(QC) procedures, or significant variation can easily occur. While Dowell performs
regular QC checks during all conductivity tests, not all laboratories do. QC checks
after an initial measurement made during the inter-laboratory study showed a
significant zero shift. Correction and comparison with a repeat test showed a 40%
variation in results. The reproducibility of the two tests after QC corrections was
within the expected variation of conductivity tests (±10 to 20% relative). Without
consistent QC and flowrate ramping to check for linearity of the pressure-drop
measurements, this type of variation would not have been discovered.
The method of pressure-drop measurement, combined with fluid flowback/flowrate
control, are changes in the Dowell-recommended conductivity test procedure that
should allow for less discrepancy between Dowell-reported conductivity data and
contract laboratory data in the future.

3.3 Other Parameters


Several factors have an as-yet unexplainable synergistic effect on the results, but
they do not show up as incremental changes when the factors are considered
individually. By necessity, these results are based on limited duplication in the matrix
of tests, with the understanding that these results will potentially form the basis for
future conductivity testing research. Optimizing these parameters to more closely
match actual reservoir/treatment conditions can result in a significant change in
conductivity results. Consistent methods of fluid preparation, direction of fluid flow,
closure and temperature profiles are part of the Dowell-recommended conductivity
test procedure.

3.4 Summary
Provided the testing procedures are consistent, it doesn’t matter where the test is
run. The results should be within 25 to 30% relative deviation. Same laboratory
testing can potentially reduce that variation to 10 to 15%. There are several key
parameters that should be checked any time conductivity results are being compared
in a stimulation treatment evaluation. The frequent major variation between data
reported by different laboratories can be reduced, provided the testing procedures
are consistent.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 1 of 32

PropNET TECHNOLOGY

1 Introductory Summary............................................................................................................. 2

2 Design ....................................................................................................................................... 6
2.1 Flowback Stability ................................................................................................................ 6
2.2 Two-Phase Flow ................................................................................................................ 10
2.3 Effective Proppant-Pack Stress Cycling............................................................................. 11
2.4 Effect of Fluid Viscosity on Proppant-Pack Stability........................................................... 12
2.5 Proppant-Pack Permeability............................................................................................... 13
2.6 PropNET Lifetime............................................................................................................... 19
2.7 Stability in Acids ................................................................................................................. 21
2.8 Effect on Proppant Settling ................................................................................................ 21
2.9 Fiber Breakage During Treatments.................................................................................... 23
2.10 Case Histories.................................................................................................................. 25

3 PropNET References ............................................................................................................. 29

4 Appendix A - Flow Test Apparatus ...................................................................................... 30


4.1 Fracture Geometry............................................................................................................. 30
4.2 Perforation Geometry......................................................................................................... 30
4.3 Tube Geometry .................................................................................................................. 30
FIGURES
Fig. 1. A molehole formed in a proppant pack of 20/40-mesh sand (FracTech test),
viewed from the 5.25 in. by 5.25 in. face. ....................................................................... 10
Fig. 2. Pressure at failure versus nitrogen flow concentration. .................................................. 11
Fig. 3. Effect of fluid viscosity on pack stability. ......................................................................... 13
Fig. 4. The effect of closure stress on proppant-pack permeability with and without J500. ....... 18
Fig. 5. The effect of J501 concentration on the permeability of 16/20 Carbolite.
(Fluid: corn syrup.) .......................................................................................................... 19
Fig. 6. Predicted proppant-pack strength with aging time for J500. ........................................... 20
Fig. 7a. Proppant settling in WF150 fluid. .................................................................................. 22
Fig. 7b. Proppant settling in WF150 fluid.
Fig. 8. The effect of J501 on proppant settling in WF175 fluid under dynamic conditions. ........ 23

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 2 of 32

Fig. 9. Resin-coated proppant containing PropNET fibers after pumping and removal
from a well. ......................................................................................................................24
Fig. 10. PropNET fibers after pumping, removal from a well and cleaning to remove oil..........25
Fig. 11. South Texas well, 15% J501 tail-in................................................................................26
Fig. 12. Fluid returns, South Texas offset wells..........................................................................27
Fig. 13. Polymer returns, offset wells. ........................................................................................28
Fig. 14. Fracture geometry test apparatus. ................................................................................31
Fig. 15. Perforation geometry test apparatus. ............................................................................32
Fig. 16. Tube geometry test apparatus. .....................................................................................32

TABLES

Table 1. Summary of Flowback and Production Rates for the Highest-Rate PropNET
Wells to date ...................................................................................................................5
Table 2. Resistance to Flowback in the Fracture Geometry.........................................................7
Table 3. Resistance to Flowback in the Perforation Geometry ....................................................8
Table 4. Failure of Proppant Packs Containing J501 in Cyclic Loading .....................................12
Table 5. Permeabilities of Various ISP and Curable RCP Proppant Packs................................14
Table 6. Permeabilities of Various Sand and Curable RCP Packs.............................................14
Table 7. Permeabilities of Precured RCP Packs ........................................................................15
Table 8. Permeabilities of 20/40-Mesh Ottawa Sand Packs.......................................................15
Table 9. Permeabilities of 12/20-Mesh Ottawa Sand Packs.......................................................16
Table 10. Permeabilities of 16/20-Mesh Carbolite Packs ...........................................................17
Table 11. Permeabilities of 20/40-Mesh Interprop Plus Packs ...................................................17

1 Introductory Summary
Propped hydraulic fracturing is successfully used in many low- and medium-
permeability formations to enhance production. A problem with some hydraulic
fractures is the back production of proppant (proppant flowback) with the oil or gas.
The proppant flowback usually occurs over a cleanup period of several days to
weeks after the fracture treatment, but can also occur throughout the economic life
of the well. Up to 20% of the proppant placed in the fracture can return during the
cleanup period (Martins et al.: SPE 24858). The proppant that flows back has a
detrimental wear effect on the production equipment and requires the use of
separators in the production line. Concern about proppant flowback can limit the
fluid flow rates during cleanup and production.
In most cases, proppant flowback does not reduce well production. It can therefore
be concluded that the fracture does not close completely as proppant is produced.
Also, the production rate can be reduced to a point where proppant is not flowed
back; therefore, fluid drag forces appear to be necessary to dislodge or carry
proppant out of the fracture. Recent modeling by Dowell and Itasca (Asgian et al.:
SPE 28510) indicates that arch formation in the pack is important. Places in the
fracture wider than 5.5 grain diameters are inherently unstable, independent of the

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 3 of 32

effective proppant stress. In these cases, fluid flow serves to sweep proppant out of
the fracture. Closure stress and fracture width have been shown to affect the
occurrence of flowback in this modeling study and experimentally (Milton-Taylor et
al.: SPE 24821).
Proppant flowback is more likely to occur with lower closure stress or wider fracture
widths. Another possible cause is that the closure stress probably varies from point
to point in the fracture between the maximum and zero. This can occur due to
uneven settling of the proppant in the fracture. The resulting stress variation can
allow proppant to be carried out of the fracture from regions of lower closure stress.
Several techniques are used to control proppant flowback; forced closure, resin
flush, curable resin-coated proppant (RCP), random fiber reinforcement, and
thermoplastic polymer strips. Forced closure is a procedure where the fracture
pressure is released beginning immediately at the end of pumping. Forced closure
is theorized to pack the proppant back towards the perforations (reverse screenout)
and allow the fracture to close quickly before the proppant has a chance to settle in
the fracture (Ely, J.W. et al.: SPE 20708, and Ely, J.W.: Stimulation Engineering
Handbook, Pennwell Publishing, 1994). It is not always effective (Martins et al.: SPE
24858). A good review of forced closure and when it may be advantageous is that of
Barree and Mukherjee (SPE 29600).
Resin flush and curable resin-coated proppants depend on a curable polymer
coating to bond the proppant grains into a solid mass. The resin flush technique
involves pumping a curable resin into the fracture at the end of the job (Minthon and
Garvin: SPE 23421). In theory, the resin coats the proppant in the fracture near the
wellbore and cures through a polymer crosslinking reaction. Additional post-flushes
are used to insure the resin does not fill the pores in the proppant pack. The
disadvantages of this technology are the difficulty in covering all of the interval with
resin, the difficulty pumping the post-flush through the entire treated volume and the
need to drill the excess resin out of the wellbore after it cures.
Curable RCPs are pre-coated with resin. They are used as all or some (tail-in) of the
proppant in the fracture. Under sufficient closure stress, shut-in time and
temperature, the resin coating cures to form a strong mass. Curable RCPs can have
any of several disadvantages. They interact with the fluid chemistry (for example,
pH value, cross-linkers and breakers). This results in additional expense due to the
necessity of adding additional chemicals, and unpredictability of curable RCP
performance. This can also result in reduced proppant-pack permeability and/or
RCP effectiveness. The resin manufacturers have worked to reduce this interaction
and some of the newer products are less reactive. The problem does require that
specific curable RCPs should be checked for reactivity with the fracturing fluid on a
case by case basis.
Curable RCPs need specific temperature, shut-in time and stress conditions to form
a strong bond. Shut-in time can be more than 24 hours, and at low temperatures
(less than 150°F [66°C]), additional chemical activators must be added to promote
cure. At high wellbore temperatures the curable RCPs can cure before the

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 4 of 32

proppants can bond together. This results in cured particles that do not adhere to
each other. Again the manufacturers have worked on developing slower reacting
curable RCPs. Anecdotal evidence is that curable RCPs can be used in wells up to
350°F [177°C]).
Screenout with the resin-coated proppant in the wellbore especially with low
temperature activators present can require the material be drilled out. Screenout is
also an issue, because unused RCP left in the sand chief/silo must be returned to
the district and may not be reusable due to resin loss during transfer steps. Curable
RCPs can also contain a significant amount of fines (mostly resin dust) that can clog
pores and reduce permeability.
Shell has shown in the laboratory that cured RCP packs are prone to failure with
cyclic well production (Davies et al.: SPE 27382). Both proppant flowback consortia
(StimLAB and FracTech) have extensively investigated at this phenomenon. The
RCP manufacturers have developed products more resistant to closure stress
cycling, and their representatives should be consulted about this issue.
PropNET* Proppant Flowback Control Additives
Proppant flowback control additives (PropNET I ADDITIVE J500 and PropNET II
ADDITIVE J501) were developed to hold the proppant in the fracture during the
production of oil or gas (or both), and to allow more flexibility in flowback design than
possible with curable RCPs. The PropNET additives work by the physical
mechanism of random fiber reinforcement; therefore, chemical curing reactions are
not necessary for the proppant to be held in place. No combinations of temperature,
pressure or shut-in time are required. PropNET has been used in over 1100
treatments with a high rate of successful proppant flowback control. Experience
indicates that wells can be flowed back at rates of thousands of B/D. Also, flowback
is possible within 15 minutes after the fracturing treatment is completed. This rapid
flowback rate allows increased polymer cleanup early in the flowback (Anderson et
al.: SPE 36468, Howard et al.: SPE 30495) and can result in increased total polymer
returned. Because no curing reactions occur, wellbore clean-out is similar to that of
normal proppant. Details of a few wells with high flowback and production rates are
provided in Table 1.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 5 of 32

Table 1. Summary of Flowback and Production Rates for the Highest-


Rate PropNET Wells to date
Flowback rates
Area Well type Overall rate Rate/perforation
WTX oil 2160 B/D 114 B/D
BCA oil 2500 B/D 50+ B/D
SET gas 5760 B/D 96 B/D
STX gas 3600 B/D 50 B/D
Production rates
WTX oil 200 B/D 10 B/D
OUK oil 9600 B/D 24 B/D
STX gas 5900 Mcf/D 92 Mcf/D
YMX gas 12000 Mcf/D 65 Mcf/D

There are several issues with PropNET fibers that need to be considered. PropNET
fibers can cause difficulty for the operator if they are returned in high levels, or the
operator is unprepared to deal with this. Therefore, PropNET addition requires
attention to assure PropNET fibers are not pumped in the flush stage. PropNET
fibers in the flush can clog chokes during flowback. Further the client needs to be
made aware of our recommendations about the use of a dual choke system for
flowback (discussed in the PropNET I Additive J500 and PropNET II Additive J501
manual section in the FRACTURING MATERIALS MANUAL — ADDITIVES).
Another difficulty with fibers during flowback is in wells with low bottomhole
pressures that require a downhole pump. If these wellbores are not cleaned out with
coiled tubing prior to production, the fibers may foul the downhole pumps.
The addition of PropNET fibers lowers the permeability of the proppant pack. The
permeability is 30% lower than a control proppant with the same fluid for 20/40-mesh
proppant. The reduction can be 50% for larger ceramic proppant (16/20 and 12/18
mesh).
A phenomenon which has been consistently observed in the laboratory and
anecdotal evidence available from the field, (Fig. 1) is that of the formation of
moleholes. Moleholes are infinite-conductivity channels that can grow from the
perforation back into the fracture by failure of the proppant/fiber pack and production
of a small amount of proppant. Once formed, these channels have been shown to
be stable in the laboratory to long-term fluid flow and to cycling of closure stress.
Formation of these moleholes may negate the decrease in conductivity due to
addition of PropNET to the proppant pack. The exact mechanism controlling the

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 6 of 32

formation of, and the benefits available from the formation of, moleholes is still
unclear, but work is underway to understand the phenomenon and quantify potential
benefits.
PropNET additives have also been used in conjunction with resin-coated proppants
to allow rapid flowback of the fracturing fluid while the resin-coated proppant is
curing. Field applications of this technology in North Sea wells has shown that the
addition of PropNET fibers to resin-coated proppant allows rapid fluid flowback with
minimal proppant production, compared to offset wells with only resin-coated
proppant. Laboratory work has also indicated that addition of PropNET to resin-
coated proppant significantly increases the resistance of RCPs to closure stress
cycling even for the most cyclic loading resistant RCPs. The laboratory work has
also shown that if failure occurs in an RCP/J501 pack, it occurs gradually with
formation of moleholes, whereas RCP packs fail suddenly. For example, under
identical conditions in recent tests an RCP pack failed suddenly after 15 stress
cycles while the same RCP with 1.5% J501 had not failed after 60 stress cycles -
only two moleholes had formed.

2 Design

2.1 Flowback Stability


A key performance criteria is proppant-pack stability when fluid is flowed through the
pack. Pack failure is defined as proppant production, and usually occurs at a
specific flow rate or pressure drop in the laboratory. Three types of tests are used to
determine the PropNET additive/proppant-pack stability to flowback. The first test
approximates flow in the fracture with the PropNET additive/proppant-pack held
between two rock faces with 1000 psi (6.9 mPa) closure stress. The second test
approximates the perforation and fracture near the wellbore with no closure stress
applied. The third geometry was a simple tube (no closure stress) with a washer
placed in the end to approximate a perforation. Drawings and discussions of the
tests are provided in Appendix A.
Pressures and flow rates for failure change from one experimental apparatus to
another. This is due to several factors; test cell geometry, closure stress, how the
cell is packed, proppant type and proppant size. None of the experiments performed
replicate field conditions and it is therefore impossible to quantify the failure point of
PropNET fibers in the field from laboratory-derived data. However, the more
different types of experiments performed, the greater the confidence in the product
performance. The suitability of a combination of these experiments in developing
effective proppant flowback control additives has been confirmed by the success of
over 1100 PropNET treatments.
Water was used as the flowback fluid in these tests. Higher viscosity fluids and
gas/water flow are discussed below.
Table 2 shows the typical laboratory pack stability to flowback encountered with
PropNET additives in the fracture geometry. In the fracture geometry, the pressure

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 7 of 32

drop is the best measure of pack strength. Table 3 shows typical data for PropNET
additives in the perforation geometry. In this geometry, flow rate is the key variable
because the pressure drop is not linear around the perforation.

Table 2. Resistance to Flowback in the Fracture Geometry


1.5% PropNET Additive Added to Proppant
Closure Stress 1000 Psi (6.9 mPa)
Values are at the Point of Pack Failure
J501 J500
Proppant 0.25-in. 0. 5-in. 0.75-in. 0. 5-in.
width(psi/ft width(psi/ft width(psi/ft width(psi/ft
[gal/min]) [gal/min]) [gal/min]) [gal/min])
Sand
20/40 mesh -- 20 [1.3] 30 [2.0] --
12/20 mesh 23 [1.6] 17 -- 6.7 [0.9]

Ceramic
20/40 mesh 6 [0.5] 7 [0.6] -- --
12/18 mesh 13 [1.9] 6 [1.2] -- --

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 8 of 32

Table 3. Resistance to Flowback in the Perforation Geometry


1.5% PropNET Additive Added to Proppant
No Closure Stress
Values are at the Point of Pack Failure
J501 J500
Proppant 0. 5-in. width 2.0-in. width 0. 5-in. width
(psi/ft (psi/ft (psi/ft
[gal/min]) [gal/min]) [gal/min])
Sand
20/40 mesh 23 [2.6] >27 [>5.1] [2.7]
12/20 mesh 15 [1.9] -- [1.5]

Carbolite
20/40 mesh 8 [1.5] >17 [>3.7] --
16/20 mesh 6 [1.8] 16 [4.4]
12/20 mesh 2 [1.1] 4 [2.9] --

Bauxite
18/30 mesh [1.6] -- --
16/20 mesh [2.3] -- --

AcFRAC Black
20/40 mesh [3.4] -- --
Several observations can be made from Table 2 and Table 3. The pack failure flow
rates are very high. One gal/min (3.8 l/min) equals 34 B/D through the perforation
(Table 3). For a 30-ft perforated zone with 120 perforations (assume only half
connect with the fracture), the 20/40-mesh sand pack containing J501 would fail at
5300 B/D (2.6 gal/min/perforation). The pack failure flow rate from the fracture
geometry is similar. The 4-in. (10 cm) wide fracture test cell has a failure flow rate of
1.3 gal/min for the 20/40-mesh sand pack containing J501. A 30-ft (9.2 m) high
fracture having the same failure flow rate over each 4 in. of its height would fail at
3980 B/D.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 9 of 32

The PropNET additive stabilizes sand, ceramic and pre-cured resin-coated proppant
packs. The pack stability to flow changes with the proppant type and size. The
PropNET additive appears less effective with sand sizes larger than 20/40-mesh,
and with ceramic proppants. Field data with 20/40-mesh ceramic proppants
(Interprop, Econoprop, Sinterball) and 12/20-mesh sand indicate that sufficient pack
stability to flow is obtained to control flowback.
J500 has not been as extensively tested as J501. Perforation geometry tests with
J500 and 20/40-mesh sand show failure flow rates equal to J501 (2.7 gal/min). J500
is very similar to J501 in dimensions and properties, and is expected to perform in a
similar manner.
The fracture width behind the perforation was also varied in the perforation geometry
(Table 3). The failure rate for J501 and 20/40-mesh sand increased from 2.6 gal/min
to greater than 5.1 gal/min (10.2 to 17 l/min) when the fracture width changed from
0.5 in. to 2.0 in. (1.3 to 3.8 cm). The same 0.5-in. (1.3 cm) perforation was used in
each test.
Several tests with J501 and 20/40 sand were performed at FracTech Ltd., Sunbury-
on-Thames, UK The test cell was loaded with a 20/40-mesh sand with 1.5% J501
slurry to produce a pack 5.25” by 5.25” by 1” with a completely open face (5.25” by
1” - no perforation) for proppant production. The pack was shut in for 12 or 24 hr at
165°F (74°C) and 500 psi for the fracturing fluid to break. The closure stress was
increased to 1000 psi for the start of flow of heated water-saturated nitrogen through
the pack. The closure stress was increased along with the flow rate to correspond to
increased closure stress with increased drawdown. The proppant/J501 pack was
able to withstand a drawdown of 40 psi/ft (1760 psi closure) across the pack without
significant proppant production. At 40 psi/ft (287 l/min gas), a channel formed
through the pack (a molehole) with approximately 28% of the pack produced
(Fig. 1). The arrows point to the edge of the molehole within the sand pack. Gas
flow was from left to right in the photograph and closure stress was applied
perpendicular to the visible face.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 10 of 32

Fig. 1. A molehole formed in a proppant pack of 20/40-mesh sand (FracTech test),


viewed from the 5.25 in. by 5.25 in. face.
The channel production caused a significant increase in pack conductivity. Following
formation of the channel, the pack stabilized and sand production decreased to
0.0001g/minute (less than 40 lbm/year for a 100 ft high fracture) at 288 l/min gas
flow rate. The gas flow rate was then increased to 573 l/min causing production of a
small amount of proppant (20g in 9 min); However, the pack stabilized at this new
flow rate and after 7 hr the sand production rate was 0.0086 g/min (7.5 lbm/day for a
100 ft high fracture).

2.2 Two-Phase Flow


Experiments were performed with two-phase gas/water flow in the tube geometry.
The pressure drop at proppant-pack failure is plotted as a function of percent volume
of gas in the fluid (Fig. 2). Gas/water is expected to be the worst case because the
surface tension difference is the highest. Note the triangle point on the graph.
Surfactant was added to the water to reduce the interfacial tension. Proppant-pack
stability increased by 20%. Water/oil surface tension is 20 to 60% that of water/air
depending on the composition, and the two-phase flow effect will be less important
than with gas/water. Note that in pure gas flow, pack strength sometimes exceeds
the maximum flow and pressure drop of the test apparatus (plus symbols). A similar
significant effect of two phase gas/water flow on the strength of resin coated
proppant packs has been observed by R.J. Vreeburg et al.: (SPE 27382) and was
attributed to higher drag forces on the proppant particles during two phase flow.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 11 of 32

Pressure at Failure, psi/ft


140
+
120
100
80 +
60
Water/N2
40
20 Pulse Flow
+ : No Failure
0 Surfactant
0 20 40 60 80 100
Nitrogen, % Total Volume Flowing

Fig. 2. Pressure at failure versus nitrogen flow concentration.


Results (Fig. 2) indicate that in the worst case, two-phase flow can lower the pack
resistance to flow by 60 percent. The design criteria for flowback rates per
perforation account for this reduction in proppant-pack resistance to flow.

2.3 Effective Proppant-Pack Stress Cycling


There is industry concern about the ability of curable resin-coated proppants to resist
cyclic stress loading (Vreeburg et al.: SPE 27382), although recent developments in
RCP technology have significantly improved their stress-cycle behavior. Cyclic
stress loading occurs when the well is shut-in and opened repeatedly. At shut-in, the
reservoir pressure is close to the in-situ stress and the effective proppant-pack stress
is low. During production, the effective proppant-pack stress is highest (in-situ stress
minus bottomhole flowing pressure), and fluid is flowing through the proppant pack.
The fluid pressure changes result in stress cycles on the proppant pack. The
PropNET additives have been tested for stability to this type of situation in the
laboratory and the field.
J501 was tested for cyclic stress loading with 20/40-mesh sand and Carbolite. The
tests were performed in a smaller size fracture geometry. The proppant packs
containing the PropNET additive were placed in a conductivity cell (7 by 1.5 in. [17.8
by 3.8 cm]). This was in turn placed in a press. The stress on the proppant pack
was cycled between 1000 and 4000 psi (6.9 and 27.6 mPa). At the higher closure
stress, water was flowed through the pack to a specific pressure drop (flow rate)
between 10 and 40 psi (69 to 276 kPa). The pressure drop (flow rate) for a specific
test was the same for every cycle. The number of cycles at which the pack failed
was recorded (Table 4). Final fracture width was 0.35 in. (9 mm).

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 12 of 32

Table 4. Failure of Proppant Packs Containing J501 in Cyclic Loading


(1000 to 4000 PSI [6.9 to 27.6 mPa])
Fluid Pressure Drop 20/40-mesh Sand 20/40-mesh Carbolite
10 psi >30 cycles >30 cycles
22 psi 14 cycles --
40 psi 15 cycles --
Table 4 shows that proppant packs containing the PropNET additive appear to be
stable to cyclic loading over several different flowing pressure drops. A field test was
performed in south Texas. A well fractured using J501 was repeatedly shut-in and
opened to determine if the proppant pack was stable. This well cleaned-up and was
producing dry gas. The in-situ stress was 9500 psi (65.5 mPa). The reservoir
pressure was 7600 psi (52.4 mPa) The effective proppant-pack stress at shut-in was
1900 psi (13.1 mPa). The flowing well head pressure was 4200 psi (29 mPa),
(effective stress 5300 psi [36.5 mPa] neglecting flowing gas friction) with 3.2
MMscf/D gas. The well was shut-in for a four-hour period and then reopened for a
two-hour period. This was done four times. At the end of the shut-in cycles the well
head pressure was 4800 psi (33.1 mPa). During the flowing periods no more than
four tablespoons of proppant were produced in a two-hour period.

2.4 Effect of Fluid Viscosity on Proppant-Pack Stability


The effect of fluid viscosity on proppant-pack stability has been examined. This is
important in cases where higher viscosity oils and/or unbroken fracturing fluid flow
through the proppant pack containing the PropNET additive. Tests were performed
in the tube geometry with two different Newtonian fluids (sucrose/water,
glycerol/water) and two different proppants (20/40-mesh sand, 16/20-mesh
Carbolite). The viscosity of the fluids was varied, ranging from 1 to 500 cp. The flow
rate at pack failure versus fluid viscosity is shown in Fig. 3.
In Fig. 3, the failure flow rate of the pack decreases with fluid viscosity to the
2/3 power. This implies that higher viscosity fluids (at bottomhole conditions) result
in lower maximum recommended flowback rates. The maximum flowback rate
equation (provided in the PropNET I Additive J500 and PropNET II Additive J501
manual section in the FRACTURING MATERIALS MANUAL — ADDITIVES)
accounts for fluid viscosity.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 13 of 32

EFFECT OF VISCOSITY ON PACK STABILITY


100

PACK FAILURE FLOWRATE (bbl/d/perf.) 20/40 SAND


SUCROSE/WATER

10

16/20 CARBOLITE
GLYCEROL/WATER
1

0.1
1 10 100 1000

FLOWBACK FLUID VISCOSITY (cP or mPa-s)

Fig. 3. Effect of fluid viscosity on pack stability.

2.5 Proppant-Pack Permeability


The permeability of proppant packs containing the PropNET additive were measured
in the Dowell conductivity laboratory. The data are shown in Tables 5 to 11.
Laboratory testing indicates proppant packs with the PropNET additive have
permeability 70% or more compared to 20/40-mesh proppant packs without fibers.
In 16/20-mesh and larger proppants, addition of PropNET additives can result in a 50
percent decrease in permeability compared to the proppant without fibers. Because
of the inert nature of the PropNET additive, the permeability results are independent
of fluid or proppant type (sand, ceramic, pre-cured resin-coated proppant). Also, the
proppant packs containing the PropNET additive have the same permeability
response to closure stress as proppant packs without the additive. Proppant packs
containing the PropNET additive can have better or worse permeability than curable
RCP packs depending on choice of proppant and fluid. For example, sand and
PropNET additive at 4000 psi (27.6 mPa) closure stress may be worse than curable
RCP because of sand crushing. Intermediate-strength proppant packs containing
the PropNET additive would have higher conductivity than curable RCP at this
closure stress.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 14 of 32

Table 5. Permeabilities of Various ISP and Curable RCP Proppant Packs


(5000 psi [34.5 mPa] closure stress)
Fluid Type Proppant J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
YF130HTD 20/40 IP none 200 250°F (121°C) 145 (fit)
Plus
YF130HTD 20/40 IP 1% 201 250°F (121°C) 106
Plus
YF130HTD 20/40 IP 1.5% 199 250°F (121°C) 114
Plus
YF140HTD SUPER HS none 208 250°F (121°C) 10
YF130HTD AcFrac none 192 250°F (121°C) 48
BLACK
Test conditions: shut-in at 3800 psi (26.2 mPa) closure for 12 hr, pressure increased to 5000
psi (34.5 mPa) for test. The fit for the control was based on 10 different points ranging from 0
to 500 lbm/1000 gal polymer loading with YF100HTD fluids and 20/40-mesh Interprop Plus.

Table 6. Permeabilities of Various Sand and Curable RCP Packs


(3000 psi [20.7 mPa] closure stress, YF120LG and Enzyme Breaker J134L)
Fluid Type Proppant PropNET Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
YF120LG Ottawa none 223 175 °F (79°C) 48
sand
YF120LG Ottawa 1.5% J500 212 175 °F (79°C) 39
sand
YF120LG Ottawa 1.5% J501 195 175 °F (79°C) 50
sand
YF120LG AcFrac CR none 247 175 °F (79°C) 42
YF120LG AcFrac SB none 282 175 °F (79°C) 49
YF120LG SUPER LC none 244 175 °F (79°C) 38

Table 6 shows a conductivity comparison between Ottawa sand with the PropNET
additives and several curable resin-coated proppants in a fluid (YF120LG)
considered to be the most compatible to curable RCPs. The fluid had an initial pH
value of 9.7, and an enzyme breaker was used. Curable RCPs are known to interact

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 15 of 32

with oxidizing breakers. Within the experimental error of the test, the PropNET
additives did not reduce the permeability of the Ottawa sand, and the sand
containing the PropNET additives and curable RCPs had equivalent permeability.

Table 7. Permeabilities of Precured RCP Packs


(5000 psi [34.5 mPa] closure stress)
Fluid J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
YF100HTD none 280 250°F (121°C) 10, 40, 41
YF100HTD 1.5% 270 250°F (121°C) 21, 30

Table 7 shows that J501 has little effect on permeability when added to a pre-cured
resin-coated proppant.

Table 8. Permeabilities of 20/40-Mesh Ottawa Sand Packs


(4000 psi [27.6 mPa] closure stress)
Fluid J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
2% KCl none 0 250°F (121°) 116
YF130HTD none 193 250°F (121°) 32
YF130HTD 0.5% 188 250°F (121°) 50
YF130HTD 1.5% 199 250°F (121°) 28
YF130HTD 2.5% 203 250°F (121°) 27
YF545HT none 205 250°F (121°) 56
YF545HT 1.5% 200 250°F (121°) 31

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 16 of 32

Table 9. Permeabilities of 12/20-Mesh Ottawa Sand Packs


(3000 psi [20.7 mPa] closure stress)
Fluid J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
2% KCl none 0 225°F (107C) 376
YF135 none 191 100°F (38°) 235
YF135 1.5% 202 100°F (38°) 157
YF340LPH none 173 225°F (107C) 98
YF340LPH 1.5% 176 225°F (107C) 72
corn syrup none 0 140°F (60°C) 488
corn syrup 1.5% 0 140°F (60°C) 346

Table 8 and Table 9 show that PropNET additive has a constant effect on
permeability regardless of proppant type, size or fluid composition. In Table 9, corn
syrup was used to show the effect of J501 on 12/20-mesh sand in fluid without
polymer. It was not possible to properly mix J501 and 12/20-mesh sand in water
containing 2% KCl.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 17 of 32

Table 10. Permeabilities of 16/20-Mesh Carbolite Packs


(4000 psi [20.7 mPa] closure stress)
Fluid J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
2% KCl none 0 250°F (121°C) 700
YF130HTD none 334 195°F (91°C) 332
YF130HTD 1.5% 269 195°F (91°C) 155
YF125LG none 279 195°F (91°C) 435
YF125LG 1.5% 227 195°F (91°C) 261
YF240D^ none 76 170°F (77°C) 246
YF240D^ 1.5% 57 170°F (77°C) 149
ClearFRAC* none 0 195°F (91°C) 686
ClearFRAC 1.5% 0 195°F (91°C) 478
^3800 psi closure stress

Table 10 shows the effect of PropNET additive on the permeability of 16/20-mesh


Carbolite proppant packs. The PropNET additive gives a decrease in permeability of
between 30% and 50% of that of the base depending on the fluid used.

Table 11. Permeabilities of 20/40-Mesh Interprop Plus Packs


(5000 psi [34.5 mPa] closure stress)
Fluid J501 Conc. Polymer Conc. Temperature Perm.
(wt/wt) (lbm/1000 gal) (Darcy)
YF130HTD none 175 250°F (121C) 164 (fit)
YF130HTD 1.5% 176 250°F (121C) 110

The effect of closure stress on Interprop Plus proppant-pack permeability is shown in


Fig. 4. The proppant pack containing J500 exhibits the same behavior as the
proppant pack without J500.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 18 of 32

800

700 20/40 INTERPROP PLUS


PERMEABILITY (DARCY)

600

500

400

300
INTERPROP PLUS WITH
1.2% PropNET I ADDITIVE J500
200

0 2000 4000 6000 8000 10000

Effective Proppant Closure Stress (psi.)

Fig. 4. The effect of closure stress on proppant-pack


permeability with and without J500.
Figure 5 shows the effect of PropNET level on the permeability of 12/20-mesh
ceramic proppant. Corn syrup was used as the mixing fluid and the tests were
performed at 4000 psi closure and 195°F (91°C). Note that reducing the PropNET
level from 1.5% to 0.5% increases permeability from 77% to 87% of the control
without PropNET.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 19 of 32

PERMEABILITY
(DARCY)
750

700

650

600

550

500

450

400
0.0 0.5 1.0 1.5 2.0 2.5

J501 CONCENTRATION

Fig. 5. The effect of J501 concentration on the permeability of 16/20 Carbolite.


(Fluid: corn syrup.)
Field Conductivity Results
A client in South Texas analyzed production-decline data to estimate fracture
conductivity from three fractures in the same well. The first was done on a fracture
containing 20/40-mesh Interprop 1. The fracture conductivity was 1550 md-ft. This
fracture produced excessive proppant, so a patch of resin-coated Interprop 1 was
pumped. The production decline indicated a fracture conductivity of 860 md-ft with
the curable RCP patch. The upper zone of this well was fractured using Interprop 1
containing J501. The production decline indicated a fracture conductivity of 1500
md-ft.

2.6 PropNET Lifetime


The PropNET materials are made of glass. Glass is a meta-stable phase and can
dissolve in water depending on temperature, time, composition of the glass, pH
value and dissolved minerals in the water. The process occurs more quickly at high
temperature, at acidic- or basic-pH values and in water without dissolved minerals.
The dissolution process is known to begin at the surface and progress inward. It is
generally a linear function of time (unless a passivating surface layer forms), but an
exponential function of temperature.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 20 of 32

During the development of the PropNET additives, the diameter of the fibers was
found to be important in determining proppant-pack stability. If the glass fibers
dissolve over time, their effective diameter and the pack stability will decrease.
Concerning the fracture, several factors are important. First, the time that the
PropNET additive is in contact with the fracturing fluid is too short to have any effect
(1 to 2 days) at less than 300°F (149°C); therefore, the water of interest is the
formation water. In most fractured sandstone formations this water is silica
saturated, has a pH value of 5.5 to 8, and has other dissolved species such as
calcium that slow the dissolution process. Finally, as time passes after the fracture
treatment, the well depletes, requiring less pack stability.
Aging experiments were performed with silica-saturated water flowing through
proppant packs containing the PropNET additive at 350°F (177°C), 300°F (149°C)
and 250°F (121°C). The tests were run for 3 months, and at various times, samples
were removed and tested for pack resistance to flow in the tube geometry
(Appendix A).
J500 followed normal dissolution behavior and the proppant-pack strength
decreased linearly with exposure time. Fig. 6 is a result of these experiments.
Percent of initial pack strength is plotted versus time for various BHSTs.

J500 - Pack Strength after Dynamic Ageing


100

90
Pack Strength, % Initial

80

70 50°C [122°F]
70°C [158°F]
90°C [194°F]
60
110°C [230°F]

50

40
0 2 4 6 8 10
Time, Years

Fig. 6. Predicted proppant-pack strength with aging time for J500.


J501 does not follow normal dissolution behavior. The proppant packs containing
J501 lost about 30% of the pack strength in two weeks at 350°F (177°C). After that,
there was no further decrease. This behavior may indicate that either an equilibrium
was reached, a protective layer formed on the surface or some other phenomenon.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 21 of 32

The sand packs containing J501 showed much better pack strength retention than
packs containing J500 at all times. J501 is expected to be more durable than J500
and have a higher use temperature.
There is always risk in extrapolating the data at 350°F (177°C) to lower temperatures
and longer times. Dissolution rate normally doubles for every 10°C increase in
temperature. At 300°F (149°C), J501/proppant-pack stability should be sufficient for
up to two years (assume 50% of the original stability). Extrapolation of stability to 10
years at lower temperatures is not recommended because of the limited time period
of the laboratory aging tests (3 months).
The first well treated in May 1994 with a reservoir temperature of 275°F (135°C) has
been sand free. A dry gas well with a reservoir temperature of 355°F (179°C) was
treated in February 1996 and has been sand free.

2.7 Stability in Acids


J500 and J501 were tested for acid stability. Samples were tested for 24 hr at 175°F
(79°C) and 1000 psi (6.9 mPa) (autoclave pressure), and 6 hr at 300°F (149°C) and
5000 psi (34.5 mPa). Samples were exposed to 15% HCl, 28% HCl, and NAS (Non-
Acid Solvent) fluids. Weight change was measured after exposure.
J500 was strongly attacked by the concentrated HCl. The J500 lost approximately
30% of its mass during the tests. The NAS had little effect on J500, causing a small
weight gain of 5%. The J501 was hardly affected by either the concentrated HCl or
the NAS. Slight weight gains approximating 4% were seen after exposure to the
fluids.
Based on these results, J500 should not be used if the well will be treated with HCl in
the future. J501 should be resistant to HCl exposure. Both PropNET additives
should be stable to NAS exposure. Neither PropNET additive will survive exposure
to treatment fluids containing HF.

2.8 Effect on Proppant Settling


Adding the PropNET additive to the proppant slurry has the added benefit of
reducing proppant settling. This can be seen in the laboratory by pouring slurries
into graduated cylinders and observing the height of the proppant pack as a function
of time. Fig. 7a and Fig. 7b show the height (volume) during settling of 20/40-mesh
sand and 16/20-mesh Carbolite in WF* 150 at room temperature.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 22 of 32

240
220
2% PropNET I J500
200
PACK VOLUME (mL)

180
160 1% PropNET I J500
140
120 NO PropNET additive
100
80
60
40
FLUID: UNCROSSLINKED 50 LB./1000 GAL. GUAR
20 INITIAL PROPPANT CONCENTRATION: 8.3 PPA
0
0 50 100 150 200 250 300 350

TIME (min)

Fig. 7a. Proppant settling in WF150 fluid.

200

2% PropNET I J500

150
PACK VOLUME (mL)

1% PropNET I J500

100 NO PropNET additive

50

FLUID: 50 LB./1000 GAL. GUAR SOL'N.


INITIAL PROPPANT CONCENTRATION: 8.3 PPA
0
0 100 200 300 400

TIME (min)

Fig. 7b. Proppant settling in WF150 fluid.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 23 of 32

Similar behavior is also seen in dynamic settling conditions. A slurry of 20/40-mesh


sand was made with WF175 at room temperature, both with and without 1.5% J501.
The slurry was placed in a transparent vessel with a cup rotating at 45 rpm in the
center. The annular gap between the rotor and vessel was 0.16-in. (0.4 cm), giving
a shear rate of around 65 sec-1 in the annular gap. The amount of free liquid was
monitored as a function of time (Fig. 8). The addition of J501 to the slurry
significantly reduces the rate of proppant settling: this is clear based on the amount
of free liquid.

Sand
6

Sand + J501
5

0
0 20 40 60 80 100 120
Time (min)

Fig. 8. The effect of J501 on proppant settling in WF175 fluid under dynamic conditions.
The PropNET additive slows the settling of the proppant and decreases the settled
pack volume: this is clear from the settling curves. This reduced settling, under both
static and dynamic conditions, should also occur in the fracture. Less settling of the
proppant is expected when the PropNET additive is used.

2.9 Fiber Breakage During Treatments


There have been concerns about the possibility of fiber breakage during mixing and
pumping operations even though yard tests have not supported this.
A fracturing treatment was performed on a North Sea well with J501 being added at
a concentration of 1.5% to a fully curable resin-coated proppant. The last proppant
stage was under-displaced and allowed to cure in the wellbore. The resin-coated
proppant containing the J501 was then drilled out. The material (shown in Fig. 9)
was circulated to a rig sand separator during the drilling operation. The fibers shown

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 24 of 32

in Fig. 9 appear to be intact, the bundles measure between 0.47 and 0.98 in.
(12 to 25 mm).

Fig. 9. Resin-coated proppant containing PropNET fibers


after pumping and removal from a well.
Fig. 10 shows fibers after cleaning to remove oil. The mass was sieved to determine
if small fiber pieces were present. The majority of the fibers were approximately
0.47-in. (12 mm) long. A few fibers were approximately 0.24-in. (6 mm) long. No
smaller fibers were found, indicating that no significant breakage of the J501 fibers
occurred during fracturing and wellbore operations.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 25 of 32

Fig. 10. PropNET fibers after pumping, removal from a well and
cleaning to remove oil.

2.10 Case Histories


South Texas
Over 140 successful fracturing treatments using the PropNET additives have been
performed in southern Texas. This case history reviews typical flowback and
polymer cleanup out of one fracture. The well has two sandstone zones separated
by a 15-ft (4.6 m) shale zone. The gross height is 100 ft (30.5 m). The net height is
25 ft (7.6 m). The depth is 8800 ft (2682 m). The average permeability of the
producing zones is 0.8 md. A 28-ft (8.5-m) section and a 8-ft (2.4-m) section were
perforated at 4 spf. The total number of perforations is 146. One large fracture with
226,000 lbm (102,514 kg) of 20/40-mesh Interprop 1, and 1730 bbl (275 m3) of
borate-crosslinked guar fluid was pumped. J501 was added to the last 15% (vol) of
the proppant.
The flowback history is shown in Fig. 11. Flowback began as soon as the treating
equipment was disconnected (30 min). The initial flowback rate was 500 B/D
(79.5 m3/D) and increased to 1000 B/D (13.7 B/D/perforation). Gas broke through
after 6 hours and the water rate decreased. The total proppant flowed back was 125
lbm (56.7 kg). This was 0.05% of the total proppant placed in the fracture. Normal
experience with the PropNET additive in southern Texas is 50 to 350 lbm (23 to
159 kg) of proppant are produced during cleanup. Most proppant is produced in the

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 26 of 32

two-phase flow time-frame when gas has broken through and before the water rate
decreases significantly. The gas rate reached 10.6 MMcf/D for this well. The gas
was flowed to sales the next morning, 12 hr after gas broke through. Total flowback
time was four days.

SOUTH TEXAS WELL, 15% PropNET J501 TAIL-IN

10000 140
WATER & GAS RATE (BBL/D, MCF/D)

CUMLATIVE PROPPANT (LBS.)


120

1000
100

80
100
PROPCUM
WATER
GAS 60

10 40

20
1 10 100

TIME (HOURS)

Fig. 11. South Texas well, 15% J501 tail-in.


The rate of fluid returned and guar polymer concentration in the fluid were evaluated
from this fracture (well A) and an offset well (well B). The offset well had two
fractures. The top zone was fractured using J501 in the last 15% (tail-in) of the total
proppant volume. The lower zone was fractured using curable resin-coated
proppant in the last 23% (tail-in) of the total proppant volume. Fluid flowback from
the three fractures is shown in Fig. 12. The two fractures containing J501 have
much higher initial flowback fluid rates and earlier gas breakthrough than the fracture
containing curable resin-coated proppant. The client’s experience was that fractures
containing curable resin-coated proppant could not be flowed back at rates faster
than 250 B/D without excessive proppant production. All three fractures returned
fluid at about the same rate after gas breakthrough.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 27 of 32

FLUID RETURNS SOUTH TEXAS OFFSET WELLS

450

CUMLATIVE FLUID RETURNED (BBLS) 400 WELL B UPPER (PROPNET)

350

300
WELL A (PROPNET)
GAS
250

200
GAS
150
WELL B LOWER (RCP)
100

50

0
0 10 20 30 40 50 60 70

FLOWBACK TIME (hours)

Fig. 12. Fluid returns, South Texas offset wells.


Samples of the flowback fluid were taken at one-hour intervals during the flowback.
The samples were analyzed for guar concentration using a chemical tagging
technique and UV/visible spectroscopy. The polymer concentration in the flowback
fluid for well A (PropNET additive) and the lower zone of well B (curable RCP) were
constant for the first two days. The polymer concentration in the flowback fluid of the
upper zone of well B steadily decreased over the first two days.
Based on the guar concentration in the flowback fluid and the rate of fluid return, the
rate of polymer return can be calculated. The fraction of polymer pumped into the
fracture that was returned is shown in Fig. 13. The two fractures containing the
PropNET additive had rapid early polymer return rates, while the fracture containing
curable RCP returned polymer gradually. This is a function of the fluid flowback rate.
The two fractures containing the PropNET additive returned a higher fraction of the
polymer in the 50 to 60 hour period; however, the fracture containing curable RCP
may have returned the same amount of polymer after longer flowback times.
Polymer return analysis from other fractures containing the PropNET additive show
as much as 50% of the polymer returned in two days.
This case history shows that the PropNET additives can be used as a tail-in to
control proppant flowback. The use of the PropNET additive allows faster flowback
rates than are commonly possible with curable resin-coated proppants, earlier gas
breakthrough and earlier gas to the sales line. In many cases, shorter flowback time
is observed, lowering cleanup costs. The PropNET additive allows for flexibility in
flowback rate, to maximize the polymer cleanup.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 28 of 32

POLYMER RETURNS OFFSET WELLS


30

WELL A ( PROPNET)
25
PERCENT POLM. RETURNED

WELL B UPPER ( PROPNET)


20 GAS

15
GAS
WELL B LOWER (RCP)

10

0
0 10 20 30 40 50 60 70

FLOWBACK TIME (hours)

Fig. 13. Polymer returns, offset wells.


Indiana
Several low-temperature gas wells were fractured in the New Albany shale of
southern Indiana. In all cases, multiple fracturing treatments were performed on a
single well. No treatment contained more than 25,000 lbm of 20/40-mesh sand, and
J501 was added to the last 50% of the proppant in most jobs.
Previously, these types of wells were fractured using curable resin-coated proppant
to control proppant flowback. This required expensive activator to be added to the
fluid. The wells were shut-in overnight to allow the resin-coated proppant to cure.
The packer was then moved, the cured resin-coated proppant in the wellbore was
drilled out, and the next fracture was performed. The total time to fracture four
zones in a well was 4 to 5 days.
Wells containing J501 were flowed back within 10 min after the end of the job. The
wells were flowed for 30 min. Very little proppant (approximately 27 lbm) was flowed
back. No drilling or clean-out was required, and the packer could be moved soon
after flowback. All four zones could be fractured in a total elapsed time of only 8 hr.
By using the PropNET additive, cost savings included reduced job costs (no
activator), and reduced rig time (approximately $1000 to $1400 USD per day).

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 29 of 32

3 PropNET References
US Patents 5,330,005, 5,439,055, and 5,501,275
SPE 30495 Howard, P. R. et al.: “Fiber/Proppant Mixtures Control Proppant
Flowback in South Texas.”
SPE 31093 Romero, J. and Féraud, J. P.: “Stability of Proppant Packs Reinforced
with Fiber for Proppant Flowback Control”.
SPE 35326 Prado-Velarde, E. et al.: “Proppant Flowback Control in the Burgos
Basin.”
SPE 36468 Anderson, A. J. et al.: “Production Enhancement Through Aggressive
Flowback Procedures in the Codell Formation.”
SPE Production and Facilities P271, November 1995, Card, R. J., et al.: “A Novel
Technology to Control Proppant Back Production.”

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 30 of 32

4 Appendix A - Flow Test Apparatus

4.1 Fracture Geometry


The fracture geometry flow test apparatus is shown in Fig. 14. The slurry is pumped
between two stone slabs and against a screen to form a pack. The screen is
removed, closure stress is applied, and water (1 cp) is pumped through the cell to
approximate production along the fracture. The flowing water drawdown (psi per
linear foot of fracture) at which the pack flows out from between the stone faces is
defined as the failure point.

4.2 Perforation Geometry


The perforation geometry flow test apparatus is shown in Fig. 15. The slurry
containing proppant and the PropNET additive is pumped through a perforation
(normally 0.5-in. diameter, but can be varied up to 1-in.) and fills a fracture. The
fracture is 4-in. high, 0.25-in. long and the width can be varied from 0.25-in. to 2.0-in.
In the fracture, the excess fluid is allowed to leakoff via a screen at the exit of the
cell, and a pack is formed. Once the pack has formed, water is flowed back through
the fracture and out the perforation. Again the pressure drop and flow rate when the
pack fails and flows out of the perforation are recorded. In this test, flowrate appears
to be the key variable. Pressure drop is more variable due to the flow patterns
around the perforation, and the shape of the arch at the perforation.

4.3 Tube Geometry


The tube geometry flow test apparatus is shown in Fig. 16. The WF130 slurry
containing proppant and the PropNET additive was pumped into the tube, and a
pack was formed against a screen placed at the exit. The screen was removed and
a washer was placed at the exit. The washer size could be changed.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Technology
Page 31 of 32

Cell

152
mm
diam
Proppant Pack

eter
Flow

100 mm long

FRACTURE GEOMETRY CELL Cell

Rock Face

Proppant Pack

Closure Stress

Fig. 14. Fracture geometry test apparatus.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Technology Dowell
Page 32 of 32

FRACTURE & PERFORATION


NO CLOSURE

F
L
U PROPPANT PACK
I
D
PERFORATION
F
1/2” to 1”
L
O
W
FRACTURE
Fig. 15. Perforation geometry test apparatus.

TUBE GEOMETRY
FLUID FLOW

PROPPANT PACK

WASHER

Fig. 16. Tube geometry test apparatus.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Flowback Information Supplement
Page 1 of 5

PropNET∗ FLOWBACK INFORMATION SUPPLEMENT

1 Wellbore Cleanup..................................................................................................................... 1

2 Flowback, Swabbing................................................................................................................ 1

3 Shut-in, Clean-up, and Production Rates .............................................................................. 2

4 Appendix................................................................................................................................... 4

FIGURES

Fig. 1. Maximum swab-cable speed vs casing ID and number of perforations............................ 4

TABLES

Table 1. Maximum Worldwide PropNET Flowback and Production Rates .................................. 5

1 Wellbore Cleanup
Prior to swabbing or flowing a well to cleanup, removal of the underdisplaced
slurry volume is recommended from the perforated wellbore area. Laboratory
and field investigations show significant tendency for fibres to separate from the
sand/liquid slurry as formation fluid or gas percolates through at low rate. The
phenomenon is similar to separation of multi-sized particles with vibration or
agitation. This may potentially result in a concentrated fibrous “mat or clump” which
can contribute to plugging of surface chokes, during the cleanup phase. In the event
excess “overhole” or “sump” is available, the slurry may be allowed to settle below
the perforations during the gel-break process. If this is not practical, removal of the
excess slurry and residual fibre from the tubing or casing may be accomplished
either through high initial flowback rates (sufficient to move and lift slurry), or
circulation of liquid or foam through coiled tubing. The latter is best completed prior
to initiating swabbing or flowback. The coiled tubing unit is also capable of
eliminating any potential proppant slurry bridges within the pipe. Bailing of sand
bridges inside pipe is not recommended due to potential pressure differentials that
can exist across the bridge.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Flowback Information Supplement Dowell
Page 2 of 5

2 Flowback, Swabbing
During flowback a collection of fibres, similar to proppant, can plug a small choke.
The use of a flowback manifold containing two parallel chokes is
recommended. A plugged choke may then be removed and cleaned while flowback
continues.
Ensure the initial wellbore volume of returned fluid is not diverted directly into
a disposal well. Initial wellbore/formation fluids should first be processed through a
separation screen as noted in HSE section above, or a 15 mesh bag-filtration unit,
where residual fibres can be separated from the fluid. If PropNET fibre is left in the
rig/return tank, the material will begin to separate from the fluid, making the fluid/tank
cleaning process more cumbersome.
PropNET flowback recommendations are based on maximum comingled fluid/water
rates. These must be measured during the swab/flowback period. Routine
strapping of the fluid-returns tank is an adequate method to determine the liquid
return rate. Foamfrac treatments present a unique situation since the water rate is
not known during the initial period when unbroken foam flows back. A five
centimeter diameter (2 in.) magnetic flowmeter and densitometer installed in the
flowback line is recommended to estimate water rate following foamfracs. Also, a
defoamer such as M45 may be added to the flowback tank in order to efficiently
gauge fluid level.

3 Shut-in, Clean-up, and Production Rates


Very short shut in times are needed for PropNET materials to control the proppant
flowback (closure on proppant). Field experiences indicate that wells can be flowed
back as quickly as 15 minutes following the end of pumping. Suggested initial
flowback rates should not exceed the maximum flowback rate calculated below.
Based on field experience, flowback stage recommendations are as follows:
• Ensure fluid viscosity is broken (i.e. water viscosity)
• Initial rate - 500 bpd or Qmax (whichever is lower) to bottoms up
• Adjust to ½ the ultimate maximum rate, then ¾, then full max. rate (each step for
2-4 hrs) (This sequence may be revised for intentional “molehole” applications for
high rate multiphase flow. See the “molehole” (i.e. infinite conductivity channel
discussion in the FEM, Section 7.9)
• Keep initial choke adjustments to < 2/64ths at a time.
As swabbing can potentially cause a flowrate in excess of Qmax, maximum swabline
pull rates are also given in Fig. 1 of the Appendix.
Dowell recommends initial cleanup and production rates not to exceed 30 bbl/day
per contributing (communicating with fracture) perforation (Qperf) for sand and
20 bbl/day per perforation for ceramic proppants. The maximum initial flowback rate
Qmax can be calculated using the equation below. Divide the Qperf by the downhole fluid

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Flowback Information Supplement
Page 3 of 5

viscosity µ2/3 for oil wells or when unbroken fracturing fluid is flowing back. The
number of perforations (Np) is divided by 2 because we assume only half the
perforations communicate with the fracture. A correction for two phase water/gas
flow has already been added to this calculation.

( )( )
Qmax = Q pert / µ 2/ 3 N p / 2

Where::
Qmax = maximum flowback rate (bbl/day)
Np = number of perforations in the fracture zone
µ = viscosity of fluid at bottom hole conditions, centipoise.
(Broken fracture fluid usually has a viscosity of 1 - 10 cp at BHST)
(Emulsified fluids will exhibit higher equivalent viscosities)
Qperf = maximum flowback rate per perforation (bbl/day/perforation)
30 bbl/day (4.8 m3/day)/perforation for sand
20 bbl/day (3.2 m3/day)/perforation for ceramic proppants

Ultimate flowback and production rates can vary significantly depending on the well,
the nature of the produced fluids and reservoir conditions. The maximum rates for
any given group of wells must be determined on a case by case basis. Qmax is
typically used as a starting point for the maximum rate, then rates are increased until
proppant is produced, or until the maximum proppant-free flowrate of the well is
achieved within production string limitations. A summary of maximum proppant-free
production rates from various areas around the world are given in Table 1 of the
Appendix.
Expected Events During High-rate Gas Flowback:
• Appearance of fiber anytime during bottoms up (dispersed or solid clumps)
• Sand and fiber strung out in wellbore, concentrated near bottomhole
• Probable sand production during/after bottoms up, typically subsides as well
cleans up
• Sand produced after each rate change, typically subsides during constant rate
• Gas break-through at 18-20% of clean fluid pumped
• 30-50% of load water returned in a few days.

Reference: PropNET Information Summary, S. D. Bittner (CDN), January 17, 1997

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
PropNET Flowback Information Supplement Dowell
Page 4 of 5

4 Appendix

MAXIMUM SWAB-CABLE SPEED VS CASING ID


AND NUMBER OF PERFORATIONS
10000
Swab Cable Speed (ft/min)

1000

100

10

1
1 2 3 4 5 6 7 8 9 10
Casing ID (in.)

5 Perforations

10 Perforations

25 Perforations

50 Perforations

ASSUMPTIONS:

1. Maximum Allowable Production Rate: 30 bbl/day/perforation. 1

2. All perforations produce at equal rates in response to swabbing.


3. All perforations produce fluid instantaneously in response to swabbing.

1 PropNET Manual Addition

Fig. 1. Maximum swab-cable speed vs casing ID and number of perforations.

DOWELL CONFIDENTIAL
Section 700.9
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
PropNET Flowback Information Supplement
Page 5 of 5

Table 1. Maximum Worldwide PropNET Flowback and Production Rates


Flowback Rates
Area Oil/Gas Overall Rate Rate/perf
(bbl/d) (bbl/d)
WTX Oil 2160 114
BCA Oil 2500 50+
SET Gas 5760 96
STX Gas 3600 50

Production Rates
Area Oil/Gas Overall Rate Rate/perf
WTX Oil 200 bbl/d 10 bbl/d
OUK* Oil 9600 bbl/d 24 bbl/d
STX Gas (wet) 5900 mcf/d 92 mcf/d
YMX Gas (wet) 12000 mcf/d 65 mcf/d
OUK* Gas (dry) 35000 mcf/d 580 mcf/d
(* PropNET w/RCP)

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 1 of 76

HyPerSTIM SERVICE

1 Introduction .............................................................................................................................. 4
1.1 Objectives ............................................................................................................................ 4
1.2 Applications.......................................................................................................................... 5
1.3 Limitations of Application ..................................................................................................... 6

2 Design ....................................................................................................................................... 6
2.1 Candidate Selection ............................................................................................................. 7
2.2 Characterization of Formation Mechanical Properties ......................................................... 9
2.3 Design Basis ...................................................................................................................... 10
2.4 Fluid Selection ................................................................................................................... 15
2.4.1 Fluid-Loss Control .................................................................................................... 16
2.4.1.1 Pressure Effects............................................................................................. 18
2.4.1.2 Temperature Effects ...................................................................................... 19
2.4.1.3 Effects of Fluid Viscosity and Polymer........................................................... 20
2.4.1.4 Effects of Fluid-Loss Additives....................................................................... 21
2.4.1.5 Fluid Selection and Fluid-Loss Control .......................................................... 21
2.4.2 The DataFRAC Service Application.......................................................................... 21
2.5 Proppant Selection and Fracture Conductivity................................................................... 22
2.5.1 Embedment .............................................................................................................. 23
2.5.1.1 Spalling .......................................................................................................... 25
2.5.1.2 Impact on Permeability .................................................................................. 26
2.5.2 Non-Darcy Flow ........................................................................................................ 27
2.5.2.1 Determination of the Inertial Flow Coefficient ................................................ 28
2.5.2.2 Non-Darcy Flow Correction of Dimensionless Fracture Conductivity ............ 32
2.5.2.3 Proppant Selection Using Manual Calculation ............................................... 33
2.5.2.4 Computer-Aided Proppant Selection ............................................................. 35
2.5.2.5 Proppant Selection Using the FracCADE Software ....................................... 37
2.5.2.6 Proppant Selection Summary ........................................................................ 37
2.5.3 Formation Sand and Fines ....................................................................................... 37

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 2 of 76

2.5.3.1 Control of Formation Fines and Sand ............................................................40


2.5.4 Proppant Flowback Control.......................................................................................40
2.6 FracCADE Software ...........................................................................................................44
2.6.1 FracNPV and QUICK Modules..................................................................................44
2.6.2 The FORECAST Module...........................................................................................47
2.6.3 The PLACEMENT II Simulator ..................................................................................49

3 Execution ................................................................................................................................54
3.1 Batch-Mix Operations .........................................................................................................54
3.2 Continuous-Mix Operations ................................................................................................55

4 Evaluation ...............................................................................................................................55
4.1 Prats’ Correlation................................................................................................................56
4.2 Modified McGuire-Sikora Correlation .................................................................................56

5 Fluid-Loss Data ......................................................................................................................64


5.1 WF120 (J164) Containing 25 lbm J478/1000 gal and 25 lbm J418/1000 gal
BHST=150°F (66°C), Pressure=1000 psi ..........................................................................64
5.2 WF160 (J164) Without Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................65
5.3 WF160 (J164) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................66
5.4 WF110 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure- 1000 psi .............................................................................................................67
5.5 WF120 (J424) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................68
5.6 WF130 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................69
5.7 WF140 (J424) Containing 50lbm J238/1000 gal  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................70
5.8 WF160 (J424) Containing Various Fluid-Loss Additives  BHST=150°F (66°C),
Pressure-1000 psi ..............................................................................................................71
5.9 YF140 (J424)  BHST=150°F (66°C), Pressure-1000 psi ................................................72
5.10 YF140 (J424)  BHST=150°F (66°C), Pressure-1000 psi ..............................................73
5.11 YF140 (J424)  BHST=175°F (79°C), Pressure-1000 psi ..............................................74

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 3 of 76

5.12 YF140HTD (J424)  BHST=200°F (93°C), Pressure-1000 psi ...................................... 75


5.13 YF140HTD (J424)  BHST=250°F (121°C), Pressure-1000 psi .................................... 76

FIGURES
Fig. 1. Production rate sensitivity to skin...................................................................................... 8
Fig. 2. IPR curve sensitivity to skin. ............................................................................................. 8
Fig. 3. The effect on shifting an 80% damage collar. ................................................................. 11
Fig. 4. Productivity-increase curves. .......................................................................................... 12
Fig. 5. Effective wellbore radius for pseudo-radial flow.............................................................. 14
Fig. 6. Fluid-loss data for YF140................................................................................................ 17
Fig. 7. Pressure gradient through a sand pack versus gas flow rate, darcy and
non-darcy flow................................................................................................................. 27
Fig. 8. Total pressure drawdown versus transit time, sanding prediction. ................................. 39
Fig. 9. Curable resin-coated proppant compressive strength required to prevent flowback. .... 44
Fig. 10. Proppant Editor. ............................................................................................................ 45
Fig. 11. FracNPV Input. ............................................................................................................. 45
Fig. 12. Equivalent wellbore radius and pseudo-skin................................................................. 47
Fig. 13. PRODUCTION FORECAST input................................................................................. 47
Fig. 14. Production simulation, non-darcy flow. ......................................................................... 49
Fig. 15. ROCK input. .................................................................................................................. 50
Fig. 16. ZONES — layer data input........................................................................................... 51
Fig. 17. PLACEMENT SIMULATOR — conventional design, 20/40-mesh sand,
1400 gal pad. ............................................................................................................... 57
Fig. 18. PLACEMENT OUTPUT — conventional design, 20/40-mesh sand. ............................ 57
Fig. 19. PLACEMENT SIMULATOR — P3D tip-screenout design, 20/40-mesh sand,
1600 gal pad. ............................................................................................................... 58
Fig. 20. PLACEMENT OUTPUT — P3D tip-screenout design, 20/40-mesh sand.................... 58
Fig. 21. PLACEMENT SIMULATOR — conventional design, 12/20-mesh sand,
1800 gal pad. ............................................................................................................... 59
Fig. 22. PLACEMENT OUTPUT — conventional design, 12/20- mesh sand. .......................... 59
Fig. 23. PLACEMENT SIMULATOR — P3D tip-screenout design, 12/20-mesh sand,
3500 gal pad. .................................................................................................................. 60
Fig. 24. PLACEMENT OUTPUT — P3D tip-screenout design, 12/20-mesh sand.................... 60
Fig. 25. Stage front propogation. ............................................................................................... 61
Fig. 26. Fracture height profile ................................................................................................... 61
Fig. 27. Wellbore fracture width profile. ..................................................................................... 62
Fig. 28. Fracture height growth history. ..................................................................................... 62
Fig. 29. Fracturing (net) pressure profile.................................................................................... 63

TABLES

Table 1. Water Viscosity at Temperature................................................................................... 20


Table 2. Dry Proppant Pack Intertial Coefficient Factors ........................................................... 31
Table 3. Proppant Selection With Embedment and Non-Darcy Flow ........................................ 35

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 4 of 76

1 Introduction
Fracturing technology has traditionally been applied to low-permeability formations to
stimulate production. More recently, fracturing has been successfully applied to
high-permeability formations (from 10 md to more than 500 md).
The practical application and rewards of fracturing high-permeability formations
include damage by-pass and stimulation through creation of a large effective
wellbore radius. The fracture treatments are designed to overcome limitations of
conventional matrix treatments. Matrix treatments often fail to provide sustained
production response because of factors such as layered or heterogeneous
permeability, acid sensitivity and formation deconsolidation. An alternate means of
providing effective wellbore communication and stimulated response is required.
Because of these factors, the HyPerSTIM* design and execution technology was
developed. The difficulty of achieving effective fractures in high-permeability
formations must be considered. In addition to achieving effective fluid-loss control,
the major considerations include generation of adequate fracture-conductivity
contrast (considering embedment and the impact of non-darcy flow), characterization
of soft formations and post-treatment control of proppant mobility.

1.1 Objectives
The HyPerSTIM Service is the fracturing design, execution and evaluation service
dedicated to fracturing moderate - to high-permeability formations.
The primary objectives of the HyPerSTIM Service include:
• to extend fracturing services to exploit high-permeability reservoirs by providing
an effective horizontal and vertical communication pathway, a pathway possibly
blocked by difficult-to-remove formation damage
• to improve the high-rate well productivity of “unstable” formations where fines
mobility resulting from large pressure drawdowns associated with radial matrix
flows in low-cohesion sands is a problem
• to complement sand-control services for unconsolidated low-cohesion sands
• to overcome the limitations of matrix acidizing and other solvent treatments when
high-permeability formations exhibit unacceptable response, or where damage is
too deep and extensive to remove (cost-effectively) using matrix injection
techniques
• to provide a means and design methodology to improve all conductivity-limited
fracture treatments.

1.2 Applications
Damage Bypass
* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 5 of 76

Fracturing (versus large matrix treatments) can be a cost-effective means of


achieving zero- or negative-skin responses. The limiting factors are fracture
conductivity and the ability to place proppant because of high fluid-leakoff rate.
Damage mechanisms can include, but are not limited to:
• cement or mud filtrate damage or both
• inadequate number of perforations or plugged perforations
• scale deposition due to pressure changes near the wellbore
• fines mobility caused by high matrix flow velocities
• low-cohesion formation collapse due to large pressure drawdown
• non-darcy flow effects and resulting pseudo-skin at high rates.
Connecting Reservoir Permeability
Laminated reservoirs with multiple vertical lithology and permeability variations can
be interconnected through a high-conductivity vertical fracture to increment reserves.
Matrix diversion may be difficult to achieve or perforating multiple zones may prove
costly. Horizontal wells drilled in large high-permeability zones may have similar
application if additional smaller laminae are present or if limited vertical permeability
exists within the drilled zone.
Mobility Control of Fines and Sand
Low-cohesion unconsolidated sands are subject to movement due to the viscous
drag imparted by the velocity of the flowing reservoir fluid. The erosion that takes
place results from exceeding the apparent cohesion stress. In addition, marginal
low-cohesion-strength formations may be destabilized and subject to fines mobility.
Pore collapse, local shear or tensile failure can occur from excessive differential
stress caused by large pressure gradients. Fracture stimulation increases the
effective wellbore radius and flow area and enables the production rate of high
permeability wells to be maintained at higher bottomhole flowing pressures, reducing
the overall effective stress applied on the rock matrix.
Stimulation
Moderate-permeability oil reservoirs (10 to 50 md) can achieve stimulation ratios
greater than two from deeper fracture penetration (100 to 300 ft) when the
CleanFRAC∗ Service technology, high-permeability proppants and fracture widening
tip-screenout techniques are used. High-permeability wells will generally be
conductivity limited, with stimulation ratios typically in the 1.5 to 2.5 range. Large
fracture widths also contribute to stimulation by reducing the flow velocity and non-
darcy inertial effects for high-rate wells.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 6 of 76

1.3 Limitations of Application


Application of the HyPerSTIM Service is considered impractical or difficult in the
following situations.
• In very-high-permeability reservoirs where high-rate fracture fluid loss restricts
proppant placement. Small-mesh proppant may be used in some cases with
restricted propped fracture-length and width.
• In reservoirs with very close gas/oil/water contacts and no fracture height-growth
restriction.
• In low-cohesion, low-compaction sands which behave inelastically (generally
greater than 30% porosity). Application of the DataFRAC* Service may be
limited. A fully packed fracture will be required, dictated primarily by gravel-
packing criteria.

2 Design
Problem definition is key in determining the application of fracturing and the
HyPerSTIM Service for high-permeability wells. Conventional techniques such as
pressure transient testing, laboratory core studies, production history evaluation, and
application of NODAL* systems analysis, can be used to quantify problem areas and
well potential. Nonmechanical, near-wellbore matrix damage can readily be
obtained through simple transient testing. Mechanical, rate, and pressure
dependent effects require more complex evaluation techniques. Additional
information regarding well performance evaluation, and discussions regarding
various skin behavior is provided in the Dowell Well Performance Manual. Examples
to be aware of include
• The determination of matrix non-darcy effects requires four-point rate and
pressure evaluation to properly assess velocity effects on pressure loss.
• Variable and deeply penetrating damage resulting from water flood alteration
of permeability (scale & salts deposition, fines mobilization) can be difficult to
determine. Core mineralogy and flow studies, combined with special transient
analysis techniques are required.
• Stress-dependent permeability requires a combination of pressure buildup
and drawdown testing.
• Matrix stability and prediction of sand movement under reservoir fluid flow and
pressure reduction requires detailed study. Previous production practices and
history will provide sufficient information for unconsolidated reservoirs.
Marginal cases will be more difficult to determine. Rock mechanics testing for
grain cohesion data and shear failure criteria may be required. Refer to
Section 2.5.3.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 7 of 76

2.1 Candidate Selection


In conjunction with problem definition, the use of NODAL systems analysis is used to
select a potential high-permeability candidate. The available software tools are the
STAR∗ software, ZODIAC* software and Systems Analysis Model (SAM+) software
for well test analysis, reservoir calibration/production prediction and NODAL analysis.
Since high-permeability reservoirs will reach pseudo-steady-state flow in a relatively
short time the applicable Darcy, Vogel/Darcy, or Jones equation should be used in
the analysis of reservoir response. The SAM manual provides additional
information. The time to reach pseudo-steady-state flow can be calculated using Eq.
1.

948 φ µ ct re2 (1)


t=
k
Where:
t = time (hr)
φ = porosity (fraction)
µ = viscosity (cp)
ct = compressibility (psi-1)
re = drainage radius (ft)
k = permeability (md).
The SAM sensitivity analyses should take into account the following factors.
1. Consider the existing well conditions to match production for various effective
skin values. Rate sensitivity to skin will provide a confirmation of damage and
an assessment of the production potential after stimulation. The fracture will be
conductivity-limited for high-permeability wells, therefore short fractures and
lower negative skin values (0 to -2) are probable. Longer fractures, with
negative skin values approaching -4, are possible for moderate permeability
wells (10 - 50 md). Fig. 1 illustrates production rate sensitivity to skin. Fig. 2
illustrates inflow performance relationship (IPR) curve sensitivity to skin.
If effective total skin is known from transient tests, the completion effects on skin
must be subtracted in order to determine the effective skin due to damage and
rate effects alone.
2. Conduct sensitivity to producing and completion effects (that is, perforation type,
shot size and density, partial penetration, gravel pack, tubing and pressure) to
eliminate potential for misapplication of the fracture treatment. A near-zero skin
may be possible by adjustment of the completion. If applicable, include the
gravel-pack completion effects after the stimulation treatment. Consider the
impact of well deviation and azimuth on postfracture productivity (Chapter 19 in
Reservoir Stimulation) when making comparisons.

∗ Mark of Schlumberger
+ Trademark of SoftSearch, Inc.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 8 of 76

3. Include the turbulent/non-darcy effects in radial matrix, gravel-pack and


perforation flows for all gas wells (Jones et al correlations) when evaluating the
impact of skin or the increase in effective wellbore radius. Very-high-rate oil
wells may also provide non-darcy effects, although for the majority of cases this
will not need to be considered.

Fig. 1. Production rate sensitivity to skin.

Fig. 2. IPR curve sensitivity to skin.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 9 of 76

2.2 Characterization of Formation Mechanical Properties


The determination of static mechanical properties (Young's modulus and Poisson's
ratio) from laboratory simulated in-situ conditions is recommended whenever
possible for initial design estimates. Effective modulus calibration through the use of
the DataFRAC Service is necessary to account for lithological variances and to
enable accurate tip-screenout designs.
Mechanical properties and acoustic wave velocities of sedimentary rock depend on
the following.
• absolute porosity (size and geometry)
• lithology (grain mineralogy and size)
• cement mineralogy and degree
• texture (compaction and grain orientation)
• type of saturating fluid.
Unfortunately, useful correlations between field dynamic data versus laboratory static
data have not been developed to enable direct use of sonic-derived properties.
Default Values
Evaluation of numerous highly porous and permeable reservoir rock data indicates
that the existing elastic modulus versus porosity correlations within the FracCADE*
software provide acceptable Young's modulus approximations in the absence of
specific data. The difficulty in using the default values when no specific data is
available, lies in proper lithologic description of the formation. The majority of high
permeability fracturing candidate sandstone reservoirs will fall into or lay between the
following three available categories.
• Quartz arenites (clean sand) — less than 10% total feldspars and clays, quartz
rich, well cemented with quartz or carbonate. Porosity correlated elastic modulus
ranges from 5.6 million psi at 10% to 2.6 million psi at 30%.
• Feldspathic/arkosic sandstones (dirty sand) — greater than 20% total feldspars
and clays, clay may be the cementing material. Porosity correlated elastic
modulus ranges from 3.5 million psi at 10% to 0.5 million psi at 30%.
• Shales/siltstones (shale) — high clay content (greater than 30%), unconsolidated
sands. Porosity correlated elastic modulus ranges from 1.3 million psi at 10% to
0.09 million psi at 30%.
For weakly consolidated, friable sandstones, the interpolated average between the
dirty sand and shale/siltstone classification is suggested. For low-cohesion,
unconsolidated sand the shale/siltstone correlation is suggested. If unsure of rock
classification between clean sand or dirty, use the average of moduli specified for

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 10 of 76

each. A review of core analyses or petrographic and SEM/XRD studies will aid in
properly classifying the rock mineralogy.
Poisson's ratio correlation is related more to the lithology and cementing material.
The value of Poisson's ratio, v, has a small impact on fracture width. Since the width
varies inversely to (1-v2)0.2 to 0.25, the average default values will be acceptable for the
majority of cases. The default value should not however, be used for stress
estimation.

2.3 Design Basis


For matrix treatments, a damage collar is removed or shifted away from the wellbore
area as illustrated in Fig. 3. The productivity improvement ratio is calculated using
Eq. 2. Practically, matrix treatments in sandstones can achieve a zero-skin effect.

(1 / ko ) log(re / rw )
Q / Qo =
(1 / ko )log( rx / rw ) + (1 / kd ) log(rc / rx ) + (1 / ko ) log(re / rc ) (2)

Where all units are consistent units, and:


Q = stimulated production rate
Qo = original production rate
ko= original undamaged reservoir permeability
kd = permeability of the damaged zone
rw = wellbore radius
re = drainage radius
rc = outer damage collar radius
rx = radius of damage removed area, inner collar radius.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 11 of 76

Fig. 3. The effect on shifting an 80% damage collar.


Assuming no completion effects, the minimum goal for the HyPerSTIM Service
application is removal of the total skin effect due to damage mechanisms.
Stimulation is possible with adequate fracture conductivity contrast. The folds of
increase in productivity of a well under pseudo-steady-state flow conditions may be
estimated using the correlations developed by Tinsely, McGuire-Sikora, Raymond-
Binder and others. As per the modified McGuire-Sikora chart, Fig. 4, the ratio of
fractured production response to undamaged radial flow is a function of the fracture
conductivity, and the fracture penetration expressed as a fraction of the drainage
radius. In most cases, for high-permeability reservoirs where fracture penetration
and conductivity ratio will be limited, the productivity index will be between 1.5 and
2.5.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 12 of 76

Fig. 4. Productivity-increase curves.


For damaged wells, similar adjusted McGuire-Sikora charts or equations of
Raymond and Binder may be used. The fractured productivity responses will
obviously be higher when compared to the damaged well productivity. Additional
discussion is provided in Reservoir Stimulation or SPE Monograph 12, Recent
Advances in Hydraulic Fracturing.
The equivalent wellbore concept proposed by Prats is the simplest to use for
estimating the effect of a finite-capacity fracture in high-permeability wells. For
fractured wells, the time to reach pseudo-radial flow (fractured case) can be
estimated using Eq. 3, when the dimensionless time is approximately equal to two
(greater than 1.6 for infinite-conductivity fractures or 2 to 5 for finite-conductivity
fractures). For high permeability, the time will be in the order of hours or days. The
higher rate produced during transient time can generally be neglected.

3,788 φ µ c t x 2f t Dxf
t= (3)
k

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 13 of 76

Where:
t = time (hr)
φ = porosity (fraction)
µ = viscosity (cp)
ct= total compressibility (psi-1)
xf = fracture half length (ft)
k = reservoir permeability (md)
tDxf = dimensionless time (approximately two).
During the pseudo-radial flow period, a fractured well behaves like an unfractured
well with an effective wellbore radius being a function of dimensionless fracture
conductivity. As the effective wellbore radius (rw′) increases, so does the production
due to a larger negative pseudo-skin effect (Eq. 4). Fig. 5 indicates that the primary
variables affecting rw′ and production are the reservoir permeability (k), fracture
length (xf) and fracture conductivity (kfw) or dimensionless conductivity (CfD).

S = − ln(rw′ / rw ), or rw′ = rw e − S (4)


Where:
S = skin effect (dimensionless)
rw' = effective wellbore radius (in.)
rw = wellbore radius (in.).
A CfD of approximately 10 or greater indicates that production is a function of fracture
length only and rw' = xf /2. For CfD values less than 10, production will be a function of
fracture length and conductivity, and rw' will be a function of the dimensionless
conductivity. The effective propped length is noticeably restricted for CfD values less
than 1.6. At very low CfD values (<0.2) production will be independent of fracture
length and will be a function of fracture conductivity and formation permeability only.
The effective wellbore radius will vary linearly with the ratio of fracture conductivity to
formation permeability and may be approximated by rw' = 0.28 kfw /k or rw' = 0.28 CfDxf.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 14 of 76

Fig. 5. Effective wellbore radius for pseudo-radial flow.


As the reservoir permeability increases, the effective propped fracture length
required for stimulation will be limited by the obtainable fracture conductivity. Once
proppant is selected, the design fracture penetration will be equal to the maximum
fracture conductivity divided by the product of formation permeability and the
minimum allowable dimensionless conductivity. Ideally a dimensionless fracture
conductivity value of 10 or greater should be created to minimize pressure losses
along the fracture length. However, even with tip-screenout designs increasing
width, this may not be practical for high-permeability cases. Other factors affecting
the fracture conductivity (polymer, embedment, non-darcy flow) are discussed in
following sections.
Prats' concept may be used to set a minimum fracture length. As previously
indicated, every effort should be made to maximize conductivity. When this is
impractical, using a suggested minimum dimensionless fracture conductivity value of
1.6 and target pseudo-skin value of -2 provides a theoretical minimum fracture
length of approximately 10 ft (for an average seven in. wellbore). In many cases, the
minimum practical fracture lengths generated and propped will be greater because
of fracture height generation, fracture compliance and minimum width requirements
prior to placing proppant. In order to generate significantly larger fracture widths and

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 15 of 76

conductivity, a longer fracture (100 to 200 ft) may be advantageous to minimize the
potential for a near-wellbore bridge and allow for longer pump times under tip-
screenout conditions.
The optimum fracture lengths will be governed by the conductivity developed and
economics. For some cases the FracNPV* module in the FracCADE software may
be used to approximate the optimum economic payout condition for very-short times
only, or when reservoir properties do not change significantly with time. When
dimensionless time is large, the analytical pseudo-steady-state flow rate
approximations coupled with decompression/depletion within the FracNPV module
do not account for changes in reservoir fluid properties. An imposed artificially large
drainage radius may be used to simulate constant fluid properties or a partial
pressure maintenance scheme. Caution must be exercised to avoid production
exceeding actual recovery capacity. If the cumulative production approximates that
obtained when using the FORECAST module in the FracCADE software for the
same time period, the simulation obtained from the FracNPV module can be
considered adequate (excepting the decline curve). Since the FracNPV module
does not simulate the effects of proppant concentration and width development due
to tip-screenout conditions, the predicted dimensionless fracture conductivity and
fracture lengths will be conservative. Sensitivity studies using very-high-proppant
concentrations can provide an indication of tip-screenout benefits for a given length.
Candidate selection sensitivity analyses may also be conducted by varying the
effective wellbore radius, assuming zero skin. Estimates of the effective wellbore
radius may be made assuming a minimum dimensionless fracture conductivity and
using Prats' correlation (previously discussed). When formation permeability is very
high, the practical limits of proppant permeability and propped width will dictate
dimensionless fracture conductivity limits.

2.4 Fluid Selection


Fracture fluid and fluid-loss additive selection is based on the following parameters.
• fluid viscosity at temperature (for proppant transport and width development)
• fluid-loss rate
• fracture conductivity damage potential
• formation damage potential.
Fracture fluid selection criteria, with respect to formation mineralogy and fluid
sensitivity, are considered similar to those used in low-permeability formation
situations. Temporary clay stabilizers and compatible chemical additives such as
scale and paraffin inhibitors may be required as part of the fracture fluid or preflush
fluid.

*Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 16 of 76

Fracturing high-permeability formations will generally require high-viscosity proppant-


transport fluids to minimize the bridging and proppant settling anticipated with the
use of larger-mesh proppant. Relatively short pump times will allow for higher
breaker concentrations to maximize pack conductivity.
Low fluid efficiency in high-permeability formations requires that emphasis be placed
on postfracture, concentrated-polymer viscosity reduction and a clean break in the
fracture and matrix. This maximizes conductivity and cleanup rate. Polymer and
particulates are filtered within the porous bed, reducing matrix permeability. Along
with deposition of solids at the fracture surface, adequate polymer buildup is
eventually developed to filter the remaining polymer within the fracture. The depth of
invasion will depend upon the ratio of both the polymer molecular size and fluid-loss
additive particle-size, to the formation pore throat-size (permeability).
High breaker concentration in fluids through the use of EB-Clean* breakers will
degrade the polymer within the fracture with time, reducing molecule size enough to
enable flowback. Concentrated polymer in the matrix may not degrade effectively
with the low breaker concentration present in most of the liquid filtrate. Based on the
results of high-conductivity fracture treatments in varying-permeability formations the
impact of matrix damage with pad fluids containing minimal soluble breaker appears
negligible due to the very large fracture-face flow area.
In general, maximizing low polymer concentration fluid viscosity through the use of
crosslinked fluids and use of CleanFRAC technology, will be preferred. Breakers
should also be added to pad fluids whenever possible. Selected fluid-loss additive
control techniques will aid in minimizing fluid loss and total polymer required.

2.4.1 Fluid-Loss Control


The fluid-loss-control mechanisms of compressibility (Cc), viscosity (Cv) and wall-
building (Cw and spurt) apply for all permeabilities. In low-permeability rock the three
regions of filter cake, filtrate invaded and reservoir fluid zones provide the pressure
differentials to control leakoff. The depth of polymer invasion is limited to near the
fracture surface and the dominant filter cake forms quickly. In high-permeability
formations, a fourth region exists where polymer and particulates invade, providing
reduced formation permeability and additional pressure loss. This is referred to as
deep-bed filtration.
Previous studies, and results of Dowell laboratory evaluation of several common
fluid/fluid-loss-additive systems, indicate leakoff to be dominated in early time by
spurt loss. The spurt volume will be a function of the square root of permeability.
Leakoff control during this period is controlled by the viscosity of the fracture fluid
and the differential pressure. Deviations from this behavior during a transition period
will depend upon the fluid particulate-to-pore diameter ratio and the rate of particle
accumulation within the matrix, which will affect the permeability. Spurt leakoff-fluid
viscosity and rheology will also change with time, due to polymer and solids filtration

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 17 of 76

within the matrix. The initial spurt losses can be orders of magnitude higher than
those for low-permeability rocks. With job pump times often less than 15 min, the
majority of leakoff can occur during the spurt and transition regions, prior to
formation of filter-cake control. Fig. 6 illustrates the time during viscosity control,
transition and filter-cake control regions for high-permeability core. The respective
fluid volume lost versus the square root of time is illustrated.

Fig. 6. Fluid-loss data for YF∗140.


The total fluid used and lost to the matrix can be reduced by reducing spurt volume.
Effective filter-cake deposition at the fracture surface will be a function of fluid-loss
additive particle size-distribution in relation to the formation pore throat-sizes,
differential pressure and the dynamic viscous-drag effects governing the transport of
fluid-loss particulates to the fracture surface. Low fluid-viscosity will aid rapid
deposition of fluid-loss additives and polymer filter cake, thereby reducing spurt time.
Using conventional matrix bridging criteria, proper sizing, in conjunction with polymer
particulates and microgels (nondispersed, hydrated gel concentrates), can minimize

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 18 of 76

the concentration of the FLA* solids required. A laboratory core evaluation for
average pore throat-size can provide guidance on selecting the particle size-
distribution for fluid-loss additives.
In contrast to low-permeability core fluid-loss data, the combination of deep-bed
filtration in high-permeability core and polymer/fluid-loss additive filter-cake effects,
provide slightly higher (similar order of magnitude) wall-building coefficient values for
conventional fluid systems. A slight decline trend in calculated wall-building
coefficient is indicated as matrix permeability (to air) increases above 50 md. Over
the test duration, this may be attributed to higher spurt velocity and the concentration
of polymer/fluid-loss additive, providing a more efficiently compacted cake, with
broader particle size-distribution. Additionally, the total pressure loss across the core
may not be dominated by the filter cake only, as is the case in low-permeability cores
with no polymer invasion. A pressure-gradient-sensitive portion of the matrix rock
will exist, which will be a function of the varying amount and depth of polymer and
fines present within the matrix. As spurt volumes increase, greater permeability
reduction from polymer and particulate filtration within the core would tend to provide
wall-building coefficient values approaching those achieved with low-permeability
core. Further detailed study is required to quantify the polymer and fines invasion
impact.
Fluid-loss data for WF and YF* fluids (illustrating the effect of additives, temperature
and pressure) are provided in Section 5.

2.4.1.1 Pressure Effects


Ideally, an instantaneous (constant time) spurt volume will be a function of the
square root of differential pressure. Assuming a constant mass of polymer and pore
bridging-material requirement, spurt volume would not vary significantly with
pressure differential. In high-permeability matrices, varying spurt time, particulate
transport, polymer invasion and filtration, changing fluid leakoff-viscosity and matrix
permeability, all result in spurt variance. Other than during the very early-time
viscosity- and compressibility-control region, the resulting interdependence of the
above variables makes a direct spurt/pressure relationship difficult to predict.
Laboratory evaluation may be best suited to approximate spurt values at pressures
different from the conventional test values of 1000 psi.
The filter cake of borate- and metal-crosslinked guar fluids (with or without fluid-loss
additives) has been determined to be nearly incompressible at pressures below 500
psi. The wall-building coefficient in this low differential-pressure range is
approximately proportional to ∆P0.6. Deviation from the ideal behavior (Cw ~ ∆P0.5) is
a result of small changes in filter-cake permeability. The filter cake is more
compressible above 500 psi, resulting in reduced cake-permeability and less
dependence upon pressure changes. Depending upon pressure range, the material-
specific, pressure exponent, varies only slightly (0.08 to 0.27). The wall-building
coefficient for the pressure ranges above 500 psi may be approximated using .
* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 19 of 76

(5)
Cw = cw 1 ( ∆P / 1,000 )0.17
Where:
Cw = wall-building coefficient at differential pressure (ft/min½)
Cw1 = wall-building coefficient at 1000 psi (ft/min½)
∆P = differential pressure (psi).

2.4.1.2 Temperature Effects


Spurt-loss volume is inversely proportional to the reduced fluid viscosity resulting
from thermal thinning and polymer degradation. Spurt-loss volume will vary
accordingly with the specific fluid system stability. The difficulty in determining the
effective shear rate and apparent fluid leakoff-viscosity for varying temperature,
permeability and differential pressure conditions, makes this correction difficult to
obtain. The system and temperature data which closely approximates the design
conditions should be used.
The wall-building coefficient variation is controlled by the viscosity of the base fluid
(for example, water). The changes will generally be small. For a given condition,
fluid system and fluid-loss additive, the wall-building coefficient may be adjusted for
temperature, if desired, using Eq. 6 and the square root of the ratio of water
viscosities at temperature provided in Table 1.

Cw 2 = Cw 1 ( µ w 1 / µ w 2 )1/ 2 (6)
Where:
Cw2 = wall-building coefficient (ft/min)½
Cw1 = wall-building coefficient (ft/min)½
µw1 = viscosity of water (cp)
µw2 = viscosity of water (cp).
The subscripts 1 and 2 refer to the respective temperature conditions.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 20 of 76

Table 1. Water Viscosity at Temperature


Temperature (°°F [°°C]) µw (cp)
75 [24] 0.91
100 [38] 0.68
125 [52] 0.53
150 [66] 0.43
175 [79] 0.36
>200 [>93] 0.30

2.4.1.3 Effects of Fluid Viscosity and Polymer


Without particulate fluid-loss control for high-permeability matrices, spurt will
increase as viscosity decreases. With particulates, the spurt time decreases as the
viscosity decreases. This is due to rapid transport of the solids to the rock fracture-
face surface, quickly establishing an effective filter cake. Additionally, the smaller
polymer molecule structure may not impede the blockage of the larger critical pore-
throats by particulates to the same degree as larger polymer structures.
The wall-building coefficient values are lowered with increasing polymer load as
more polymer residue is available to bridge with the fluid-loss additive within the
fixed time period of the test. For wall-building fluids, the polymer concentration will
dominate the leakoff coefficient. A minimum of 20 lbm/1000 gal of guar or
hydroxypropylguar (HPG) is required to provide adequate leakoff control when fluid-
loss additives are used. Studies have shown for equivalent polymer concentration,
borate- and metal-crosslinked guar-derivative fluids have similar wall-building
coefficient values, provided the metal-crosslinked systems are delayed and not
shear-degraded prior to filtration. Shear-degraded systems have substantially higher
leakoff values.
Fluids containing low-residue hydroxyethylcellulose (HEC) polymers without fluid-
loss additives are assumed to exhibit viscosity-controlled leakoff behavior. The
formation of very small microgels (nondispersed, partially-hydrated polymer)
resulting from inadequate shear in mixing, often results in gradual polymer filtration
and the formation of a filter cake. Additional shear and filtration (common in sand-
control operations) results in a cleaner fluid with higher leakoff rate, approaching true
viscosity control. Viscosity-controlled leakoff in low-permeability reservoirs (<10 md)
may prove adequate. Higher permeability will generally require the use of fluid-loss
additives, microgels and polymer residues to enable proppant placement. Data
(provided in ) indicates spurt values one order of magnitude higher with HEC.
Reasonable fluid-loss control to 50 md, may be obtained with the combined use of
graded silica Fluid-Loss Additive J84 or Fluid-Loss Additive J418 in conjunction with
FLA J478 Slurriable/Degradable Additive. Beyond 50 md the use of linear HEC

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 21 of 76

polymers should be limited. Alternative fluid systems such as the PERMPAC* solids-
free system will also have similar problems in controlling fluid loss.

2.4.1.4 Effects of Fluid-Loss Additives


Conventional silica (J418) and Diverting Agent J238 (oil-soluble resin) provide
adequate spurt and leakoff control for all permeability ranges evaluated with guar-
base fluids. J84, a larger particle size-distribution silica flour, will provide slightly
better control than J418 in larger pore sizes. To reduce the potential for fracture
conductivity damage, combinations of J418 and J238, or J418 and J478 may be
used. Neither J478 or diesel when used individually as fluid-loss agents are effective
in high-permeability formations. The combination of J478 and diesel does, however,
provide effective leakoff control with borate-crosslinked fluids and minimizes the
potential damage to fracture conductivity.

2.4.1.5 Fluid Selection and Fluid-Loss Control


Fluid-loss control (viscosity and fluid-loss additives) will dramatically affect volume
loss and proppant scheduling particularly for tip-screenout designs. An optimum
balance of viscosity, polymer load, fluid-loss additive type and concentration can be
formulated to minimize fluid loss and damage. The following methodology may be
used.
1. Select formation-compatible fluid systems for evaluation. Minimize polymer
content and use borate-crosslinked fluids when possible in conjunction with the
CleanFRAC service. In small treatments, because of spurt domination, the
slurry fluid may require a similar concentration of fluid-loss additives as the pad
fluid.
2. Apply the appropriate fluid proppant-pack damage factors.
3. Refer to Section 5 for fluid-loss data. Assume that delayed metal-crosslinked
fluid-loss data is similar to borate-crosslinked fluid-loss data for equivalent
polymer concentration. If required, apply appropriate corrections to the wall-
building coefficient and spurt.
4. Determine the most efficient or cost-effective fluid using conventional design
techniques and the FracCADE software.
5. Extend to tip-screenout design.
Additional discussion regarding fracturing fluids, additives and fluid-loss control is
provided in the respective FRACTURING ENGINEERING MANUAL sections.

2.4.2 The DataFRAC Service Application


Placement of larger, high-conductivity proppants and pumping a fracture treatment
under a tip-screenout condition requires a high degree of certainty of fracture

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 22 of 76

mechanics and leakoff control. Many high-permeability formations have lower


Young's modulus values that are often difficult to obtain through core evaluation
because of incompetent core. In the case of shales and siltstones or high-porosity,
unconsolidated sands, dynamic moduli and Poisson's ratio values are difficult to
extract from sonic data, unless specialized dipole sonic tools are used. The
Schlumberger Dipole Shear Sonic Imager (DSI*) tool uses unidirectional excitation to
extract the shear-wave-transit time from “slow” formations. Other concerns include
detection and elimination of perforation entry problems, essential for high proppant-
concentrations. Fluid efficiencies can be one-quarter to one-half of commonly
expected values for low-permeability formations.
Calibration for formation compliance and treatment efficiency is essential for all tip-
screenout designs. Since fluid loss in high-permeability formations is spurt
dominated, conducting a calibration treatment (similarly sized to the fracture
treatment) can serve two purposes.
1. To provide spurt control for the subsequent fracture treatment.
2. To obtain the effective wall-building coefficient. Note that the spurt and transition
time frame can vary (in some cases up to 10 min), prior to establishing constant
leakoff rate. This may be a large percentage of, or, as long as the fracture
treatment, and can affect the results of the DataFRAC analysis. Spurt volumes
cannot be obtained from the DataFRAC analysis, however, the spurt losses may
be disregarded if a calibration treatment is performed using fluids establishing
effective spurt control. The fracture treatment is then designed using the
combined leakoff coefficients as the effective wall-building coefficient (with zero
spurt).
The techniques described in the DataFRAC Service section of the FRACTURING
ENGINEERING MANUAL should be used. In-situ stress determination in high-fluid-
loss formations may be difficult to obtain by the pump-in/flowback technique due to
rapid leakoff. The pump-in/shut-in technique, combined with the step-rate test may
be preferred for these cases prior to the calibration treatment. If necessary,
following pad fluid injection and shut in for two or more times the pump time, a
follow-up, step-rate injection and flowback may be conducted. The spurt losses will
have been completed and a filter-cake control system established. In addition, the
poroelastic stress changes in the reservoir will be minimized. Bottomhole pressure
measurement for stress and fracture calibration is recommended.

2.5 Proppant Selection and Fracture Conductivity


The dimensionless fracture conductivity will be dominated by high formation-
permeability and effective fracture length. When compared to conventional fracture
treatment applications, reservoir permeability is two to three orders of magnitude
higher, while the fracture length is generally reduced to less than one order of
magnitude. This results from a conductivity-limited fracture. Practically, the largest

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 23 of 76

impact that can be made is to increase proppant permeability and propped fracture
width.
A number of factors influence proppant permeability and ultimate fracture
conductivity. Factors such as proppant size-distribution, time and pressure-
dependent stress changes, fracture fluid polymer residue damage, fines invasion or
scale deposition, multiphase flow and relative-permeability effects, proppant
concentration and similar effects are discussed in other literature and in other
sections of this document.
High flow rates, low-cohesion- and low-elastic-modulus (soft) sands, are often
associated with high-permeability reservoirs. Proppant embedment, non-darcy flow
and formation fines or grain movement directly impact the numerator in the
dimensionless conductivity equation, and warrant special consideration in the
HyPerSTIM Service design. Unconsolidated sands and low-cohesion rock are
subject to collapse and movement with pressure drawdown and high-velocity fluid
drag. Embedment of proppant into uncompacted, low-cohesion, low-modulus rock
significantly reduces fracture width and permeability. The high-velocity, non-darcy
flow effect imparts an additional pressure gradient along the fracture, reducing the
effective dimensionless fracture conductivity.

2.5.1 Embedment
The current data base in the FracCADE software contains undamaged, long-term,
proppant permeability data. Well-cemented Ohio sandstone core used to evaluate
proppant conductivity exhibits negligible proppant embedment. Consequently, the
high-porosity and high-permeability region adjacent to the proppant and core
interface provides a significant contribution to overall proppant permeability values.
As polymer filter cake is introduced at the wall, the porosity in this region is reduced
and the high permeability disappears, along with effective flow width. The use of an
encapsulated breaker provides a means to mitigate the effect of damage caused by
polymer residue in the pack and at the formation wall. Conductivity modification in
the FracCADE software is currently handled through an adjustment in permeability
only, using an input fluid damage factor. No allowance is made for changing flow
width due to embedment. Similarly, the effect of increasing stress on proppant
porosity and corresponding width reduction, is also handled through permeability
adjustment. The computed proppant concentration (lbm/ft2) and fixed proppant
porosity, is used to derive the effective propped width using Eq. 7.

C p = 5.2γ (1 − φ p )w (7)
Where:
Cp = proppant concentration (lbm/ft2)
γ = proppant specific gravity (dimensionless)
φp = proppant pack porosity (fraction)
w = propped fracture width (in.).

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 24 of 76

In high-permeability reservoirs, large flow rates introduce non-darcy pressure losses


that are proportional to the square of the flow velocity. The velocity is a function of
effective flow width and height. Additional proppant concentrations beyond the
minimum required to overcome embedment are required to provide additional
fracture conductivity in order to reduce any non-darcy effects.
Fracture-width loss will result from closure-stress application and proppant-porosity
loss. Additional width loss through proppant embedment results from rock failure
and compaction of formation material. The rock lithology, cementation and shear
strength will govern the degree of proppant embedment. Other contributing factors
include applied stress, proppant particle size-distribution relative to formation grains
and displacement of failed formation material into the pack. Although absolute
embedment will be less with smaller-mesh proppants, the permeability gained by
using larger mesh sizes will ultimately result in greater fracture conductivity for the
majority of cases if sufficient concentration is used. In all cases, the embedment
depth can be related to the largest or mean grain diameter of the proppant. In a
given formation, the percentage of average grain diameter lost will be approximately
the same for all mesh sizes. The maximum expected embedment per side will be
0.5 times the largest proppant grain diameter, or approximately 0.75 times the
average grain diameter, making the maximum fracture width loss 1.5 times the
average proppant grain diameter.
Correlations have been developed by the Stim-LAB Consortium relating total width
loss expressed in proppant grain diameters, to formation Young's modulus at
stresses above 2000 psi. The additional embedment effect into polymer cakes is not
included in these correlations. Although modulus does not infer formation failure and
compaction criteria, it provides a qualitative indicator of rock strength. It is often
more readily available than shear strength data (Mohr/Coulomb failure envelopes).
Generally, porous reservoir rock will have decreasing shear and embedment
strength with decreasing modulus. Specific embedment testing should be conducted
when practical.
For closure pressures greater than 2000 psi, Eq. 8 may be used to approximate
width-loss due to proppant embedment in the absence of specific data. Eq. 7 can
then be used to compute the additional proppant concentration required.

we = D p [ 0.8128 + ( −0.4191 ln E )] (8)


Where:
we = embedment-width loss, both sides, for stress greater than 2000 psi,
maximum we = 1.5 Dp (in.)
Dp = proppant average grain diameter (in.)
E = Young's modulus (million psi).
For closure pressures less than 2000 psi, a factor of 0.5 times we is suggested. The
above correlation over-predicts embedment for low-cohesion and low-modulus rock.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 25 of 76

A maximum embedment correction factor of 1.5 average grain diameters (0.75


diameters per side) is recommended for unconsolidated sands and low rock moduli.
Example:
Young's modulus (psi) = 1,000,000
Proppant size (mesh range) = 20/40
Proppant specific gravity = 2.65
Proppant porosity (%) = 35
Proppant average grain diameter (in.) = 0.025
Design goal is 2 lbm/ft2 proppant concentration in the fracture.
1. Determine the total grain diameters lost to embedment (Eq. 8).
we = Dp [0.8128 + (-0.4191 ln E)]
we = 0.025 {0.8128 + [-0.4191 ln (1)]}
we = 0.0203
2. Determine the additional required proppant concentration (above the design
goal). Use Eq. 7.
Cp= 5.2 γ(1-φp)w
Cp (5.2)(2.65)(1-0.35)(0.0203)
Cp = .18
A 2.2 lbm/ft2 proppant concentration is required to achieve the design goal.

2.5.1.1 Spalling
In very-soft, unconsolidated sands, additional width-loss, due to spalling of the
formation grains near the fracture surface, has been reported by the Stim-LAB
Consortium. Once failed and released, formation material may penetrate the
porosity of the proppant-pack if formation grains and material aggregates are small
enough.
The spalling effect reported may be considered inconclusive and an artifact of
testing. Embedment testing requires uniform normal loading, and uniform values of
shear stress at the contact faces and within the proppant-pack. The API conductivity
cell used for the Stim-LAB Consortium testing does not provide the triaxial loading
condition, due to lateral shear stresses in the proppant and core platens at the
boundary conditions of the cell. A triaxial load, conductivity test cell is better suited
to evaluate the spalling and overall embedment effect. Numerous proppant
embedment tests by Dowell were conducted (triaxially loaded formation) with
consolidated, low elastic modulus rock (0.9E6 - 3.0E6 psi) with no evidence of
spalling effects. Additional similar testing is required to confirm the results with low-
cohesion, unconsolidated formation material.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 26 of 76

In triaxially confined rock, the proppant is unlikely to continue to embed and displace
spalled material unless the low-cohesion formation rock is fluidized from drawdown.
In addition, assuming sand-control bridging criteria, all released aggregate diameters
would be required to be approximately less than one-sixth the average proppant
grain diameter.
If significant formation spalling is considered possible, the Stim-LAB Consortium
correlation (Eq. 9), developed using unconsolidated formation sand (E < 0.5E6 psi),
may be used to estimate the spalling width-reduction. The result would be added to
the embedment previously calculated using Eq. 8. The correlation is provided for
informational purposes and should be used with caution.

ws = 0.3[ we / 2D p )] 2. 5 (9)
Where:
ws = width loss due to spalling effects (in.)
we = embedment-width loss, both sides, for stress greater than 2000 psi,
maximum we = 1.5 Dp (in.)
Dp = proppant average grain diameter (in.).
Using the previous example the total width reduction (we + ws) is 0.052 in. The
proppant concentration lost to width-reduction effects is now 0.47 lbm/ft2.
Conversely, the effective width, w′, at originally designed 2 lbm/ft2 concentration is
now reduced to 0.17 in. (1.53 lbm/ft2).

2.5.1.2 Impact on Permeability


Fracture permeability will vary with embedment depth, stress and fines invasion into
the pack. When no polymer is used, permeability values 30% to 40% of
nonembedded proppant have been reported from conductivity tests with 1.5 grain
diameters embedment, severe spalling and fines mobility. Other evaluations with 0.5
grain diameters embedded and limited fines, provided 80 to 90% of nonembedded
values. Inadequate data exists for generalization. The proppant permeability at the
wall is already impaired with polymer filter cake. When the damage factor from the
polymer is considered, additional permeability correction due to embedment, should
not be required.
The wall effect is diminished as proppant concentration increases. The HyPerSTIM
Service requires high proppant concentrations for adequate stimulation. The
proppant database in the FracCADE software should be used without permeability
adjustment on the PROPPANT screen. The design proppant concentration in the
fracture should be increased by the embedment amount.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 27 of 76

2.5.2 Non-Darcy Flow


Fluid flow through porous media may be divided into three flow regimes — viscous,
visco-inertial and turbulent. The majority of flow provided by propped hydraulic
fractures can be characterized by viscous laminar (darcy) or the transitional (non-
darcy) regime. True turbulent flow is unlikely. Energy loss from non-darcy flow is a
result of fluid accelerations and decelerations through flow channels of varying size
and direction. In addition, as a result of non-streamline flow, interfering secondary
flow patterns are established above some critical rate. The additional resistance is
termed inertial flow resistance.
The non-darcy flow effect is present within porous media at very low Reynolds
numbers (greater than 0.1, less than 600). Detection is difficult until the Reynolds
number is greater than 1. Non-darcy flow exists for the majority of fractured gas
wells, but may not significantly affect production response unless fracture
conductivity is very low or rates are very high. Reservoir simulation, incorporating
non-darcy fracture flow, must be conducted to adequately predict the impact. Fig. 7
illustrates the pressure losses that may be incurred flowing gas through a sand-
packed fracture considering darcy and non-darcy flow. The viscosity of crude in oil
wells causes the viscous pressure losses to dominate for the majority of oil well
cases. High permeability, and liquid flow rate in the order of 1000 BOPD per foot of
interval may begin to provide additional non-darcy pressure losses.

Fig. 7. Pressure gradient through a sand pack versus gas flow rate,
darcy and non-darcy flow.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 28 of 76

The pressure loss for flow through proppant packs can be calculated using the
Forcheimer flow equation (Eq. 10). The first term represents the viscous effect and
the second term the non-darcy effects. The non-darcy pressure loss is proportional
to the fluid density, flow velocity squared and the coefficient of inertial flow
resistance.

∆P / L = ( µν / k p ) + βρν 2 (10)
Where:
∆P = differential pressure (psi)
L = effective length of proppant pack (ft)
µ= fluid viscosity (cp)
v = flow velocity (ft/min)
kp = proppant permeability (md)
β = inertial flow coefficient (dimensionless)
ρ = fluid density (lbm/gal).

2.5.2.1 Determination of the Inertial Flow Coefficient


Beta (β) is empirically derived from conductivity experiments for various proppants
by plotting the fluid mass flow velocity “x” (ρv/µ) versus “y” (∆P/Lµv or 1/kp). In effect,
β is the slope of the reciprocal of the permeability versus velocity relationship. A
number of factors influence the beta coefficient for proppants. These are:
• Particle size-distribution and porosity (pore to throat diameter ratio)Large-mesh
proppants with small variance in grain size-distribution results in large pore to
throat diameter ratios and inertia changes. The result at a given permeability is
an increase in β. As permeability of proppant increases (as is the case for larger
prop), β decreases.
• Grain sphericity and surface relative roughness Angular sands contribute less to
streamline flow.
• Heterogeneities such as fines, immobile oil and water saturations, and polymer
residueDebris formed and porosity reduction during closure-stress application,
results in an increase in the β factor. Cyclic stress loading of proppant causes
redistribution of grains and generated fines resulting in an increase in β. Polymer
residue also causes an increase. As fracture cleanup occurs, the polymer effect
is diminished and may actually decrease the β value due to streamlining of pore
throats. Immobile-fluid saturation contributes to reduced porosity and increases β
by two to three fold. Geertsma, Evans and Maloney provide corrections as a
function of effective porosity and residual saturation, proportional to 1/(1-Sw)5.5.
These effects are proppant and environment specific, and cannot be generalized.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 29 of 76

• Flow velocity or Reynolds numberFlow velocities, or “x” values less than 5,


provide β factors several times higher than those obtained at higher (x less than
20) rates. β values should be determined for specific rate application.
• Flowing fluid physical propertiesMulti-phase flow, relative-permeability changes
and altered flow paths have a pronounced effect on the inertial coefficient.
Corrections from dry gas values of seven- to ten-fold have been reported. During
the post-fracture clean-up period the non-darcy impact can be substantial until
the immobile liquid saturations are reduced below 10% of proppant porosity.
Numerous investigations have been conducted to estimate β for various proppants
and flow conditions. Unfortunately, inconsistency in test conditions have provided a
wide variance in available data. Cooke developed a correlation technique for β as a
function of proppant size and permeability at stress. This has become an industry
accepted approach. Unfortunately the specific proppant data developed by Cooke
was at very-low flow rates for Brady silica sands and to closure pressures exceeding
grain failure regions. The inertial flow coefficient data developed by the Stim-LAB
Consortium (for x < 20) is currently the most comprehensive data set available. The
data is incorporated into the FracCADE proppant data base for use in the
FORECAST and calculated FracNPV modules.
The inertial flow coefficient may be calculated using Eq. 11.
β = b / k ap
or log (β) = log (b ) − ( a ) log ( k p ) (11)
Where:
β = inertial coefficient (atm-sec2/gm)
kp = proppant permeability (darcies)
a = slope of log β versus log kp data line
b = intercept of log-log data line.
Refer to Table 2 for coefficients “a” and “b” for dry, proppant packs. These values
(for x < 20) are defaulted into the proppant data base in the FracCADE software.
Refer to section 2.6.1 regarding the use of the “a” and “b” coefficients in the FracIPR
software.
Corrections for proppant pack-fluid saturations or other effects are not recommended
when implementing non-darcy flow calculations unless reasonable estimates of
immobile fluid saturation are available.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 30 of 76

In the absence of proppant a and b values, the Janicek and Katz equation (Eq. 12)
may be used to approximate β for effective proppant porosity (at stress, with or
without residual fluid saturation).

(
β = 0.03167 / k p 1. 25φ e 0.75 ) (12)
Where:
β = inertial coefficient (atm-sec2/gm)
kp = proppant permeability (darcies)
φe = effective proppant porosity (fraction).

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 31 of 76

Table 2. Dry Proppant Pack Intertial Coefficient Factors


Low (X<20) and High (X>20) Rates
Beta = b/ka
Low Rate High Rate
Proppant Size a b b
Arizona Sand 12/20 1.16 0.28
20/40 1.47 0.95
40/70 2.06 6.63
Colorado Sand 10/20 1.02 0.19
16/30 1.16 0.24
16/40 1.26 0.47
Hickory Sand 8/16 1.14 0.40
12/20 1.13 0.32
16/30 1.16 0.27
20/40 1.54 1.23
Jordan Sand 12/20 0.75 0.02
16/30 0.93 0.06
20/40 1.45 0.75
40/70 1.68 0.77
1
Carbo-Lite 12/20 1.05 0.25 0.14
16/30 1.13 0.32 0.16
20/40 1.25 0.70 0.30
2
LWP+ 16/20 1.17 0.44 0.19
20/40 1.13 0.38
2
Interprop+ 12/20 0.59 0.012
16/30 0.61 0.013
20/40 0.66 0.015
Sintered Bauxite 20/40 0.98 0.10 0.06
3
Super HS 16/30 0.93 0.14 0.08
20/40 1.02 0.18 0.07
4
AcFrac PR 20/40 1.30 0.40
4
AcFrac CR 16/30 0.63 0.04 0.03
20/40 0.63 0.04 0.03
1
Trademark of Carbo Ceramics, Incorporated
2
Trademark of Norton Alcoa Proppants
3
Trademark of Santrol Products
4
Trademark of Acme Resin Corporation

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 32 of 76

2.5.2.2 Non-Darcy Flow Correction of Dimensionless Fracture Conductivity


The effective fracture conductivity will be reduced as a result of non-darcy flow.
Gidley and others have shown the reduction to be a function of a dimensionless
Reynolds number, NRe, given by Eq. 13 and Eq. 14. Taking into account the flow
distribution along the fracture, the effective conductivity, CfD′, can be expressed by
Eq. 15. Since total flow rate at the wellbore is used in the Reynolds number
calculation, a fracture flow adjustment must be applied to eliminate over-correcting
the non-darcy effect using maximum wellbore rate. Obtained from numerical
simulation, the flow coefficient, c, varies with fracture geometry and flow condition
and is generally ranges from 0.31 to 0.6. A value of 0.31 is recommended for most
fracture conditions.
βk f gq
N Re = 0.002594
µw ′h p
( gas well ) (13)

βk f γ o q o
N Re = 0.011885
µw ′h p
( oil well) (14)

C fD ′ = C fd / ( 1 + cN Re ) (15)

Where:
NRe = Reynolds Number correction (dimensionless)
β = proppant non-darcy Beta factor (atm-sec2/gm)
kf = effective proppant fracture permeability (darcies)
g = gas gravity relative to air (dimensionless)
q = gas flow rate at standard conditions (Mcf/D)
µ = average reservoir oil or gas viscosity (cp)
w′ = effective average fracture width (in.)
hp = gross producing or flowing height (ft) (the lesser of gross perforated
interval or propped fracture height)
γ°= oil specific gravity relative to water (dimensionless)
q°= oil flow rate at reservoir conditions (BOPD
CfD = fracture conductivity ratio for darcy flow (dimensionless)
CfD′ + effective fracture conductivity ratio for non-darcy flow (dimensionless)
c = fracture flow coefficient (0.31).

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 33 of 76

To determine the combined impact of proppant imbedment, non-darcy flow and


selected proppants, the FracNPV module in the FracCADE software may be used or
a series of pseudo-steady-state flow computations may be calculated using Eq. 16
and Eq. 17.

q=
(
0.000703kh Pe2 − Pwf2 ) ( gas well) (16)
(
µZ ( T + 460) ln 0.6[ re / rw′ ] )
Qo =
(
0.00708 kh Pe − Pwf ) ( oil well) (17)
(
µBo ln o.6[ re / rw′ ])
Where:
q = gas flow rate (Mcf/D)
qo = oil flow rate (BOPD)
µ = average reservoir oil or gas viscosity (cp)
k = reservoir permeability (md)
h = net height (ft)
Pe = average reservoir static pressure (psi)
Pwf = wellbore sandface flowing pressure (psi)
Z = average super-compressibility factor (dimensionless)
T = reservoir temperature (°F)
re = drainage radius (ft)
rw′ = effective wellbore radius (ft)
Bo = formation oil volume factor (bbl/STB).

2.5.2.3 Proppant Selection Using Manual Calculation


At a selected flow rate, the additional proppant concentration required to
compensate for the combined effects may be calculated. Different proppants can be
compared to determine the most cost-effective proppant. The following example
illustrates the computations involved.
Example:
Proppant
− 20/40-mesh sand, 0.025 in. avg. grain size, 35% porosity, 2.65 specific gravity
− a = 1.45, b = 0.75 (from Table 2)
− kp = 140 darcies at closure stress
− kf = 98 darcies with 0.7 fluid damage factor

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 34 of 76

Rock Modulus
− 1,000,000 psi
Oil Reservoir
− k = 100 md
− Pe = 1350 psi Pwf = 700 psi
− re = 1050 ft rw = 0.54
− µo = 0.8 cp
− γo = 0.81
− h and hp = 30 ft
Design goal
− S = -2, CfD = 1.6
Steps
1. Compute the effective wellbore radius (Eq. 4), rw′ = 4.0 ft.
2. Determine the minimum required fracture length (Eq. 5), xf = 15 ft.
3. Determine pseudo-steady rate (Eq. 17), qo = 3411 BOPD.
4. Compute the first effective fracture width required from CfD, w′ = 0.29 in.
5. Estimate the width loss due to embedment (Eq. 8), we = 0.02 in.
6. Compute the total proppant required (for w′ + we) to achieve w (Eq. 7),
Cp = 2.8 lbm/ft2.
7. Determine β, (Eq. 11), β = 0.00058 atm-sec2/gm.
8. Compute the first estimate of Reynolds Number (Eq. 14), NRe = 0.4.
9. Determine the non-darcy effective CfD′, (Eq. 15), CfD′; = 1.4.
10. Make adjustment for non-darcy effect, increase w′ and Cp by (CfD /CfD′)0.52,
w′ = 0.31 in., Cp = 3.0 lbm/ft2 (the alternative is an increase proppant permeability).
11. Recompute NRe and CfD′ to determine the effect of increased width.
In most cases, one or two iterations will provide a dimensionless fracture conductivity
close to the design requirement. For the above example, the practical minimum
fracture length should be increased (to approximately 50 ft or more) to allow for a tip-
screenout design to achieve the high proppant concentrations required. This will
result in the requirement to reduce practical design dimensionless fracture
conductivity. Alternatively, a higher permeability proppant may be selected using
similar methodology to increase conductivity at the reduced width and proppant
concentration.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 35 of 76

In practice, when dealing with high-permeability oil wells, the correction for non-darcy
flow will be insignificant. The additional width required to achieve reasonable darcy
conductivity will dominate, as illustrated in the previous example.

2.5.2.4 Computer-Aided Proppant Selection


A spreadsheet-base algorithm, (PROPPIC in the 20/20+ software) has been
developed using the methodology presented in Proppant Selection Using Manual
Calculation and may be used in lieu of manual calculation. The 20/20 spreadsheet,
PROPPIC, has the following features.
• Using various proppants, equivalent propped widths required to achieve equal
production are determined (equivalent CfD′).
• Comparisons are shown to determine the proppant that will result in the least
pressure drop (largest CfD′), the least cost of an equivalent CfD′, and the least
proppant concentration (risk or difficulty).
Example input/output is illustrated in Table 3 for a gas well. The assumptions
include constant proppant porosity of 35%, constant wellbore flow velocity
throughout fracture length and all production flows through the fracture.

Table 3. Proppant Selection With Embedment and Non-Darcy Flow

+ Copyright of Access Technology, Inc.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 36 of 76

User input is available from the FracCADE software database or other conventional
sources:
Input
• Reservoir Data: permeability (md), Z-factor, bottomhole static and flowing
pressure (psi), reservoir temperature (°F), gas gravity, spacing (acres), reservoir
fluid viscosity (cp), Young's modulus (MMpsi), net height (ft), porosity (fracture),
compressibility (psi-1).
• Fracture Data: effective propped-to-net height ratio, length (ft), maximum and
minimum allowable prop concentration for normal or tip-screenout designs
(lbm/ft2), design dimensionless fracture conductivity, Reynolds number correction
coefficient.
• Column 1 — proppant description
• Column 2 — proppant permeability in darcies (used for β and conductivity)
• Column 3 — non-darcy “a” exponent for β calculation
• Column 4 — non-darcy “b” intercept value for β calculation
• Column 5 — proppant mean diameter (in.)
• Column 6 — proppant specific gravity (gm/cc)
• Column 7 — proppant cost ($/lbm)
Output
• Time to pseudo-steady-state condition (day)
• Time to pseudo-steady-state (dimensionless)
• Wellbore radius (ft)
• Wellbore ratio (dimensionless)
• Equivalent wellbore radius (ft)
• Column 8 — non-darcy β factor (atm-sec2/gm)
• Column 9 — Reynolds number for embedment corrected width, minimum
proppant concentration
• Column 10 — fracture conductivity for darcy flow, md-ft
• Column 11 — uncorrected darcy flow dimensionless fracture conductivity
• Column 12 — corrected effective dimensionless fracture conductivity for non-
darcy flow
• Column 13 — conductivity multiplier required to achieve design dimensionless
fracture conductivity
• Column 14 — width loss (embedment) based on loss factor and diameter (lbm/ft2)

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 37 of 76

• Column 15 — first approximation of proppant concentration (lbm/ft2)


• Column 16 — new apparent dimensionless fracture conductivity (based on
Column 15)
• Column 17 — cost/ft2 based on conventional proppant concentration maximum.
XLIM is displayed when lbm/ft2 limit is exceeded.
• Column 18 — cost/ft2 based on tip-screenout proppant concentration maximum

2.5.2.5 Proppant Selection Using the FracCADE Software


The FracNPV module incorporates non-darcy flow effects for gas wells only.
Constant rate and pseudo-state flow are not simulated. The pseudo-steady state
flow (combined with depletion) approximations used for high-permeability wells (at
large dimensionless time) are not entirely valid for long term. The relative effects
can be considered acceptable for evaluating the non-darcy impact on proppants. In
general, the preferred order of proppants selected for conventional treatment design
will be maintained for tip-screenout, high-conductivity designs.
An illustration of modifications and use of the FracNPV module is presented in
Section 2.6, FracCADE Software.

2.5.2.6 Proppant Selection Summary


• Determine the proppant mesh size and type to consider based on conventional
operating criteria of stress, fines or proppant flowback control.
• Determine the embedment width loss and the corresponding additional proppant
concentration required.
• Estimate the effect of non-darcy flow for gas wells using computer-aided or
manual calculations.
• Select the practical, economic combination of proppant permeability and fracture
width required to minimize the effects of embedment and non-darcy flow for the
specified fracture-length target.

2.5.3 Formation Sand and Fines


Formation fines and sand production can be classified into three categories.
1. Transient sand production declines with time under constant production.
2. Continuous sand production results from unconsolidated formations.
3. Catastrophic sand production increases with time and often results from high
flow velocities and pore collapse due to pressure drawdown.
Four reservoir rock failure mechanisms lead to sand production.
1. Pore collapse occurs when effective stress in the reservoir exceeds the strength
of the rock matrix. A decrease in pore pressure and resulting gradual

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 38 of 76

compaction can result in matrix permeability reduction. Large pressure


gradients and associated increase in drag force aggravate sanding problems.
With violent pore collapse, induced macro-fractures will release loose particles.
2. Shear failure occurs when differential effective stresses become too large and
exceed the Mohr-Coulomb (shear versus uniaxial compressive strength) failure
envelope for the rock.
3. Tensile failure occurs when the tensile strength of the rock material is exceeded
by large pressure gradients. This is mostly confined to clay-cemented, low-
cohesion rock material. Reduction of the near wellbore pressure gradient
through stimulation, prior to initiating sanding, will reduce or delay sanding
potential.
4. Erosion results when loose particles are present and the fluid drag forces
exceed the cohesion stress value.
Existing sand production prediction techniques are based on field observation of
production, laboratory experimentation or theoretical modeling. A combination of
techniques may be required to assist prediction. These techniques are —
• Field derived techniques proposed by Tixier have established a limit value for a
sonic and density-log derived parameter given by Eq. 18. When the ratio is
exceeded, no sanding problems are expected. The relationship is geology and
area dependent.
G / c > 0.8 ⋅ 10 12 psi 2 (18)
b
Where:
G = dynamic shear modulus (psi)
cb = bulk compressibility (psi-1).
Another approach uses total drawdown pressure (depletion included) versus
compressional sonic transit time. Monitoring of produced debris for wells at various
drawdown pressure is required. Fig. 8 illustrates an example that has been
established using data from various fields. The position of the “risk” region is field
dependent. The data, in absence of other specific data, may be used with caution as
an indication of sanding risk.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 39 of 76

Fig. 8. Total pressure drawdown versus transit time, sanding prediction.


• Laboratory experimentation using a triaxially loaded core to determine the failure
envelope (linear cohesion stress value and angle of internal friction) with varying
pore pressure will provide the relationship between the effective stress, shear
failure and critical pore pressure. Thick-walled hollow cylinder tests will aid in
determining the critical pressure gradient at which sanding problems are initiated.
• The Schlumberger IMPACT* data processing system may be used for rock
mechanics evaluation and to predict sanding tendencies. The data processing
system uses shear and compressional wave transit times (derived from the
Dipole Shear Sonic Imager), density, pore pressure, minimum in-situ stress,
laboratory rock strength parameters, porosity and geochemical logs, and other
data to forecast theoretical critical pressure and sanding tendencies.
• Theoretical modeling requires a mathematical formulation of the sand failure
mechanism, often difficult to obtain. Validation with laboratory and field sand
production data is essential. Without calibration, current theoretical tools are
primarily for qualitative use.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 40 of 76

2.5.3.1 Control of Formation Fines and Sand


Current sand-control technology involves the use of sized gravel and screens to limit
sand movement and permeability reduction near the wellbore. High-conductivity
fracturing offers one additional mechanism for marginal sand producers; reduction of
pressure gradient and fluid velocity at the same producing rate through effective
wellbore enlargement.
Low-Cohesion, Unconsolidated Reservoirs
• Control is required through follow-up gravel packing (The STIMPAC Service).
Average grain diameter of the gravel (proppant) should be six times the average
grain diameter of the formation sand at the cumulative 50th percentile weight.
This criteria should be considered for wells with a previous history of continuous
sand production where the probability of disrupted sand aggregates and
discretized particles near the wellbore exists.
• The gravel-pack design and gravel and screen selection criteria are provided in
the Sand Control Engineering Handbook. Conventional sand-control hardware
may be used for small slurry volumes when fracturing at low rates (less than
15 bbl/min) with the tools in place. For higher-rate treatments the use of a
modified and hardened crossover assembly is recommended (see Section 2.5.4,
Proppant Flowback Control).
• Fracture proppant selection may be modified and grain size increased if
laboratory core/proppant flow evaluations are available. Average grain size to
18 times the average grain diameter at the cumulative 50th percentile by weight
has been shown to be effective due to formation grain agglomeration and lower
pressure gradients.
Marginal Sand Producers, Friable-Consolidated Reservoirs
• Determine the minimum effective stress level where sand production occurs.
• Conduct production rate sensitivity to generated fracture geometry with various
proppants at maximum allowable drawdown (less safety factor). Compare to
radial flow rate obtained for an undamaged matrix at equivalent differential
pressure.
• Select proppant using previously discussed methodologies.

2.5.4 Proppant Flowback Control


Factors for determining excessive proppant flowback in high-permeability wells
include combinations of the following.
• Low closure stress — Low closure stress limits the normal force applied to the
proppant grains.
• The proppant grain properties of sphericity, relative roughness and surface
friction coefficient; — Angular proppants tend to bridge, causing them to

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 41 of 76

withstand higher drag forces prior to movement. Because of permeability loss,


however, angular proppants are limited to low-stress environments.
• Grain size-distribution and packing — Similar sized particles provide higher
permeability but are less likely to bridge.
• Proppant density — Untrapped proppant grains of higher density are less likely to
be mobilized by fluid moving through the fracture than lighter grains.
• Proppant concentration in the fracture — Monolayers, or layers less than 3 grain
diameter widths, are more likely to contain proppants. They do however, provide
lower conductivity and higher relative embedment and non-darcy effects.
• Effective drag forces acting on the proppant grains — High flow velocities create
larger dynamic pressure and fluid momentum forces. Non-darcy flow increases
the fluid drag and effective friction factor. These effects are magnified by
perforating only a portion of the expected fracture height.
• Effective flowing fluid viscosity — High-viscosity fluids or multiphase flow
introduces large effective viscosity resulting in increased drag.
The above combined effects for proppant packs have not been fully investigated and
modeled. Given the fracture requirements of the HyPerSTIM Service, two major
techniques of proppant flowback control are available — mechanical isolation using
the STIMPAC Service proppant exclusion system or the use of curable resin-coated
proppant.
Proppant Exclusion Using Mechanical Devices
Isolation packers, wire-wrapped screen hardware and modified crossover tools are
used to conduct the fracture treatment through and enable control of severe
proppant-flowback and fines production. The basic equipment list includes the
following.
• Model 18 sump packer — The Model 18 sump packer locates and supports the
screen, blank pipe and Model 21 production packer. The Model 18 allows
passage of debris into the rat hole.
• Model 18 seal assembly — The Model 18 seal assembly allows communication
through the Model 18 sump packer.
• Long-stroke PosiTest* packer — In deviated (greater than 60 degrees) wells
where wireline setting is not possible, the PosiTest packer can be used.
• A screen or wire-wrapped perforated tubing — A screen or wire-wrapped
perforated tubing sized for fracturing proppant may be used. A prepacked screen
may be used as an alternate. Five to ten feet of overlap above and below the
under-reamed casing window is recommended. A SUPER-SET+ screen handling
table should be used to minimize screen damage.

* Mark of Schlumberger
+ Trademark of Johnson Screen

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 42 of 76

• Blank pipe or spacer tubing — A blank pipe or spacer tubing is used below the
Model 21 packer to allow for extended seal assembly and tubing movement
compensation. A blank pipe or spacer tubing is also used below the safety shear
collar for separation of the crossover assembly from screen. The length normally
equals the screen length.
• Safety shear collar; — The safety shear collar provides a means of releasing the
Model 21 packer and circulating housing enabling retrieval from the well.
• Crossover assembly — The crossover assembly allows for high-rate (greater
than 50 bbl/min) slurry placement behind the screen with minimum erosion.
Following placement, the port-isolation assembly is closed and the shift assembly
retrieved using coiled tubing, allowing production communication through the
screen.
• Model 21 packer — The Model 21 packer is a retrievable production packer set
hydraulically or on wireline. The differential pressure rating is 6000 psi at 210°F
(99°C).
• Model 21 seal assembly — The Model 21 seal assembly provides for tubing
connection to packer. Connection is via a no-go locator, collet locator or snap-
latch locator. A collet locator assembly is preferred for fracturing operations to
ensure positive location and allow seal movement.
• Coiled tubing fishing/shifting tools — These are overshot tools with circulation
subs and knuckle joints. Following coiled tubing circulation of proppant debris
above the fishing neck, the tools are used for shifting crossover assembly ports
and for retrieval of the shift assembly.
The STIMPAC Proppant Exclusion System manual provides basic operation,
procedures and parts information.
Curable Resin-Coated Proppants
The use of curable phenolic resins with proppants, has long been recognized as a
method for controlling proppant flowback. A tail-in treatment technique using curable
resin-coated proppant for 10% to 20% of the total proppant pumped, is used in the
majority of treatments. The technique has met with mixed results.
Norman, et al conducted laboratory evaluations using curable resin-coated proppant
to determine the strength required to prevent flowback. Fig. 9 may be used to
approximate the required minimum compressive strength of a cured pack relative to
the production rate through a 0.5 in. perforation. Refer to SPE 20640 for additional
information.
When curable resin-coated proppants are considered, design considerations should
include the following.
• The potential of premature proppant bridging at high proppant concentrations
may preclude total coverage of perforations by the curable resin-coated
proppant. This may result in proppant flowback from some intervals. Multi-layer

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 43 of 76

perforated intervals, vertical wellbores with greater than 20 ft of gross


perforations and highly deviated (greater than 45°) wellbores are candidates for
the use of curable resin-coated proppant throughout the slurry, as opposed to
curable resin-coated proppant in just the tail-in slurry.
• Incompatibility with high-pH (greater than 10) borate-crosslinked fracture fluids.
Curable resin-coated proppant reduces the pH value 1 to 1.5 units. The result is
less available crosslinker and reduced fluid viscosity. Additional crosslinker (2 to
3 times) may be added to compensate. In the case of nondelayed systems,
additional activator may be added.
• Incompatibility with oxidizing breakers, Breaker J218, J475, or J479. The type of
curable resin-coated proppant, weight percent of coating, free resin dust and
proppant concentration will determine the threshold amount of ammonium
persulphate required to effect a break. Once the threshold amount is exceeded,
the required breaker will be proportional to polymer concentration. Two to ten
times as much breaker may be required to overcome the neutralization effect of
the resin.
• Compressive and tensile strength of the curable resin-coated proppant is reduced
with high pH values and oxidizing breaker concentration. A minimum
compressive strength of 150 psi is recommended for high-rate wells to minimize
potential flowback problems. When curing acceleration is required in crosslinked
YF fluids, 3 to 5% (vol/vol) Isopropyl Alcohol F3 is preferred.
Specific fluid/temperature/time laboratory studies are recommended for all high-pH
fluids containing oxidizing breakers or alcohol stabilizers. Fluid performance and
proppant cured-strengths must be described for the application.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 44 of 76

Fig. 9. Curable resin-coated proppant compressive strength required


to prevent flowback.

2.6 FracCADE Software


The FracCADE options pertaining to the HyPerSTIM Service and tip-screenout
design are described in the following sections. A summary design procedure and
example is provided to illustrate the design technique using the PLACEMENT II
simulator. Detailed discussion of other module changes and additions such as MLF
(Multilayer Fracturing Model) is provided in the FracCADE User's Manual, release
notes and on-line help screens.

2.6.1 FracNPV and QUICK Modules


Refer to Fig. 10, Proppant Editor screen, and Fig. 11, FracNPV screens for the
changes.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 45 of 76

Fig. 10. Proppant Editor.

Fig. 11. FracNPV Input.


Proppant Editor
The inertial flow coefficient (Beta factor) is computed using Eq 11. Turbulence
Constants “a” and “b” provided in the proppant data base are used when β is in atm-
sec2/gm units. Proppant permeability is in darcy units.
NOTE— The FracIPR software uses β in ft-1 units and proppant permeability in md
units, hence the constants are not interchangeable. The exponent “a” is numerically
identical for both FracNPV and FracIPR input. The constant “b” required for FracIPR
input may be obtained by multiplying the FracNPV equivalent by 1000 raised to the
“a” power (bFracIPR = bFracNPV X 1000a).
Placement and Reservoir Factors
Modifications to the FracNPV and QUICK modules include the addition of two “safety
factors” to account for the analytical fracture model limitations, uncertainties of
fracture design, proppant placement and apparent producing length. The design
length is reduced by the Placement Factor to adjust for the effective propped length
relative to hydraulic length. This accounts for increased fluid loss to fissures, height
growth and reduced proppant transport from gravity current. The apparent fracture
length for production purposes is obtained from the Reservoir Factor, which is the

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 46 of 76

fraction of the propped length. The Reservoir Factor accounts for nonideal reservoir
behavior such as layered, anisotropic and stress-sensitive permeability. Respective
default values are 0.7 and 0.6 for the placement and reservoir factors. These may
be modified based on local experience. Additional information is provided in Nolte,
K.G., and Economides, M.J.: “Fracture Design and Validation with Uncertainty and
Model Limitations,”JPT (Sept 1991) 1147-1155.
Include Turbulence Effect (Yes/No)
The option for inclusion of the non-darcy effect is provided for gas wells only, and
should be used to approximate the effect on production for all high-rate gas wells.
The equations used to correct dimensionless fracture conductivity are discussed in
Section 2.5.2. The assumptions and limitations of the non-darcy correction to
dimensionless fracture conductivity include the following.
• Non-darcy flow exists within the propped fracture only. All flow is through the
fracture. It should be noted that non-darcy flow may also exist in the matrix of
wells with very short effective wellbore radii and limited matrix damage.
• Flow rate (velocity) at the well bore is used and considered constant for the entire
fracture length. The Reynolds number coefficient (c), given in Eq. 15 (to correct
for flow geometry and rate variance) is 0.31. The FracNPV module does iterate
on flow rate, and adjusts the Reynolds number appropriately with time, thereby
reducing the velocity effect.
• Gross fracture height and average fracture width is used to compute the effective
flow area and Reynolds number. The FracNPV module assumes the entire gross
fracture height is propped. The impact of non-darcy flow is therefore under
estimated if gross fracture height is larger than either gross formation interval or
propped height. If net interval height is used, the result is an over-estimation of
non-darcy effects when the propped gross formation thickness is greater than net
thickness (multi-layer zones).
FracNPV Output
An additional output page (Fig. 12) has been provided in the engineering report
displaying the predicted equivalent wellbore radius and pseudo-skin for each fracture
length and time. The information may be used for construction of decline curves
using the SAM software or for evaluation comparison to actual skin obtained. When
the dimensionless time, tDxf, is less than 10, the equivalent wellbore radius is
computed from rw′ = re /(e[PD+0.75]) where PD is the type-curve dimensionless pressure
function at each dimensionless time. When the dimensionless time is greater than
10, the equivalent wellbore radius for pseudo-radial flow is calculated from the Prats
and Cinco-Ley/Samaniego correlation (represented in Fig. 5). The pseudo-skin is
obtained using Eq. 4.
The data will be of value primarily for low- and moderate-permeability wells during
the transient flow period. The output will have limited use for high-permeability wells.
As the dimensionless time and pressure become large (most high-permeability

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 47 of 76

cases) the effective wellbore radius will decrease and pseudo-skin will increase
during the transient period.

Fig. 12. Equivalent wellbore radius and pseudo-skin.

2.6.2 The FORECAST Module


Fig. 13 illustrates the FORECAST module screen changes.

Fig. 13. PRODUCTION FORECAST input.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 48 of 76

The FORECAST module should be used to estimate production response for high-
permeability wells, within the assumed constraints of boundary conditions and drive
mechanism. The non-darcy flow effect has been incorporated within the production
forecast calculation for gas wells. The “Include Turbulence Effect” flag is defaulted
to YES for gas wells. If set to NO, the Turbulence Constant values from the
proppant data base are not displayed. The fields are also blank when an oil well is
specified.
The FORECAST module computes the non-darcy flow effect using a correction to
fracture capacity described by Holditch, which is similar to the previously discussed
Gidley correction. The differences in computation procedure and output compared
to the FracNPV computation procedure and output are described below.
• Non-darcy flow is considered within the fracture only. The fracture is divided into
cells and the flow through each cell is treated individually. The flow velocity and
Reynolds number is computed for each cell cross-sectional area, and the
correction is applied to the conductivity of the respective cell. The average
cumulative velocity is used to correct successive cells approaching the wellbore.
The numerical approach results in reduced non-darcy effects within the fracture
at length. As with the FracNPV computation, the β factor is computed from
Eq. 11.
• The gross fracture height (assumed continuous and propped) and respective
fracture cell cross section is used to determine the flow velocity. This is
consistent with the FracNPV analysis and may result in overcorrection as
previously discussed.
Example output illustrated in Fig. 14 compares FracNPV and FORECAST
computations for the same high-permeability gas well described (Fig. 11 and
Fig. 13), with equivalent fracture parameters and the following propped interval
heights.
Gross fracture height — 120 ft
Gross formation height — 94 ft
Net formation height — 50 ft

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 49 of 76

Fig. 14. Production simulation, non-darcy flow.


The procedures described in Section 2.5.2 should be used to minimize the impact of
non-darcy flow.

2.6.3 The PLACEMENT II Simulator


A pseudo-three dimensional (P3D) model has been incorporated into the
PLACEMENT numerical fracture simulator. The simulator is now capable of
modeling fracture growth into layers above and below the pay interval. The
simulator will also model fracture extension or recession. The latter capability is
important in the fracture design for high-permeability formations to model wide
fractures created by tip-screenout conditions, enabling pumping simulation after
proppant bridging. In addition, the short fractures often created for skin bypass may
be simulated through the use of a new “lateral coupling” option (P3D_LAT). This
represents a gradual evolution from KGD to radial to PKN pressure behavior. A brief
overview of the changes and features of the PLACEMENT II simulator follows.
Discussion of the new multilayer and fracture acidizing modules in the FracCADE
software are not covered. The FracCADE User's Manual release notes and on-line
help screens provide additional information.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 50 of 76

General Modifications
A PERFORATIONS screen has been added in the FracCADE software to allow for
definition of multiple-perforation intervals required by the multilayer fracture module.
Previous reference to perforations on the WELL screen has been deleted.
Nonoverlapping perforated intervals must be entered in order of increasing depth.
The RESERVOIR screen has been modified to reflect produced reservoir fluid
properties only. PVT correlation data is provided in a new ZONES screen.
ROCK Screen
The ROCK screen (Fig. 15) in the FracCADE software has been changed. The
screen defines the various layer rock data for a maximum of 10 rock types. Tip-
screenout design requires accurate data from core evaluation. As previously
discussed, the defaults for Young's modulus are lithology and porosity dependent.
The compressibility is a function of input porosity only. The default values should be
used with caution.

Fig. 15. ROCK input.


ZONES Screen
The ZONES Screen (Fig. 16) has been introduced to define the stress profile and
variation of rock properties with depth. A maximum of 20 contiguous zones may be
defined. The left portion of the screen displays the detailed properties of the
individual zone indexed. Only one zone may be defined for 2D simulations (all
modules but MLF and PLACEMENT P3D), however, if more than one zone is
defined, then the 2D modules will use the properties of the zones defined in the field
entitled Zones for 2D Calculations From (also present on other screens). Refer to
the user release notes for details on how the equivalent 2D properties are
determined and screen functions.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 51 of 76

Fig. 16. ZONES — layer data input.


PLACEMENT Screen
Fig. 17 illustrates the changes to the PLACEMENT SIMULATOR screen. The leakoff
coefficient and spurt fields were added to support the USER leakoff model
(discussed below). The Model Poroelasticity (Y/N) field allows the P3D model to
include this effect on pressure and width. The field entitled Temp. Model allows a
choice between a constant bilinear temperature profile throughout the treatment or a
general transient numerical solution to heat flow in the fracture. This option is
available in the PLACEMENT II simulator only. The fields entitled Top Barrier Flag
and Bot(tom) Barrier Flag specify the action to be taken during a P3D PLACEMENT
II simulation if the fracture height reaches the top or bottom of the defined zones.
The barrier zones are assumed to extend to infinity to allow continued height growth
or fluid loss is assumed high enough to restrict any further height growth.
The following describes the assumptions and limitations of the P3D model in the
FracDADE software.
• Individual cross sections of the fracture act independently. The pressure/width
relation is independent of length and is only a function of rock mechanical
property and fracture height. If the P3D_LAT lateral coupling option is invoked,
the pressure at any cross section will depend on the pressure everywhere in the
fracture (sections do not act independently). The correction is small for long
fractures, however, the correction may be significant in the case of short
fractures. The P3D_LAT option should be used where fracture height-to-half-
length ratio is greater than 1.0 or for simulation of early-time pressure data.
• Fracture toughness (fracture tip-effect) is considered.
• Fluid flow is horizontal. The pressure loss in the vertical direction is due only to
hydrostatic gradient. Flow into the fracture occurs over the full height. Large

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 52 of 76

relative fluid losses (greater than 60%) to low-stress porous intervals above the
pay interval are not accurately modeled because vertical flow is ignored.
• Layers immediately above and below the pay interval must have higher stress
levels than the pay interval. The overall profile entered must be able to contain
the fracture (upper and lowermost intervals must act as barriers). The thickness
of any barrier is also limited by the stress contrast between the barrier and the
pay zone.
• A constant- or variable-leakoff model with variable-leakoff rates into different
zones is supported in the FracCADE software. A USER input leakoff coefficient
may be used as an alternate. The USER specified value is assumed to be a
single, constant-leakoff coefficient for all fluids everywhere within the fracture.
• Clean (no proppant) fracture-fluid properties may be specified as constant or
varying.
• Two screenout modes, bridging and dehydration, are possible with continued
pumping. The fluid is assumed to flow through the point where proppant has
bridged. When a screenout occurs, the fracture stops growing in height and
length and the proppant concentration increases. Bridging may occur at some
point far from the fracture tip.
• The P3D model uses proppant-bridging criteria based on the average diameter of
the proppant and average cross-sectional fracture width (approximately three-
fourths of the maximum width). Along with the effect of poroelasticity, the total
result may produce proppant bridging earlier than experienced with an equivalent
single-layer 2D design.
Fracture Tip-Screenout Design
With the treatment candidate, fracture fluid and proppant selected using methods
previously described, the following example well data is used to illustrate the use of
the PLACEMENT II simulator.
Example Design Parameters—
Permeability, porosity and rock parameters per Fig. 15 and Fig. 16
Bottomhole static temperature, 200°F (93°C)
Fluid — YF140D
Ct (from DataFRAC Analysis), 0.005 ft/min½
Fracture half length is approximately 100 ft
Dimensionless fracture conductivity greater than 1.5, maximize within constraints
Maximum allowable pressure increase after screenout is 1500 psi
Maximum pump rate, 10 bbl/min

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 53 of 76

Procedure
1. Use a manual calculation or computer-aided software (PROPPIC or FracCADE) to
aid proppant selection. The FracNPV module is used to aid selection of proppant
and fracture length target for this example. (The spacing was increased to
maximum to approximate limited depletion. Caution must be exercised to ensure
cumulative production does not exceed recoverable reserves within actual spacing
limit.)
2. Use appropriate 2D equivalent geometry and a pump schedule routine such as the
INVERSE module to design a conventional treatment. This will establish an initial
trial pump schedule for the penetration and maximum prop concentration
specified.
3. When considering first design proposals without DataFRAC Service input, the
P3D model may be used initially to estimate equivalent 2D geometry and to
calibrate the Placement Factor.
4. Select P3D or P3D_LAT geometry and rerun the PLACEMENT module, increasing
the volume of high-concentration slurry in the last one or two stages. In the
majority of cases, increased pad volume will be required to prevent premature
bridging, because of potential height-growth and the slight differences in bridging
criteria between the 2D and P3D models. Intermediate stages may be adjusted
(through the INVERSE module fluid correction and proppant reduction factors) if
required for uniformity or tapering of propped fracture concentration.
Discussion of Results
The effect of non-darcy flow for this oil well was insignificant and was not considered.
Proppant embedment effect was minimal with an estimated 0.12 or 0.21 lbm/ft2
additional concentration required for 20/40- and 12/20-mesh proppant respectively.
Proppant permeability and concentration sensitivity studies indicated benefits by
using 12/20-mesh sand at maximum concentration (xf = 100 to 150 ft, avg CfD = 1.5)
instead of 20/40-mesh sand (50 to 100 ft, avg CfD = 0.5). For the example 12 md
well, stimulation is possible (S = -3 to -4).
Two options are reviewed: (1) using 20/40-mesh proppant and increasing the width
to achieve equivalent 12/20-mesh CfD of 1.5, and (2) using 12/20-mesh proppant,
further improving dimensionless fracture conductivity and incrementing production.
Fig. 17 and Fig. 18 illustrate the conventional design pump schedule and geometry
using 20/40-mesh proppant. Fig. 19 and Fig. 20 illustrate the revised schedule and
the resulting increase in dimensionless fracture conductivity, propped concentration
and net pressure.
When considering the larger 12/20-mesh proppant, additional pad volume is required
for both the conventional 2D and P3D screenout-design to eliminate early bridging.
Fig. 21 and Fig. 22 provide the conventional schedule and created geometry.
Fig. 23 and Fig. 24 provide the screenout schedule and created geometry. For
reasons previously discussed (poroelasticity and bridging criteria), the P3D

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 54 of 76

simulation indicated premature bridging of the 12/20 proppant at 64 ft. In practice,


assuming accuracy of the simulator bridging criteria at 2.5 average grain diameters,
improving fluid efficiency through rate increase may provide the only recourse to
enable use of 12/20-mesh proppant for this case. Rate and proppant transport
limitations may preclude exclusive use of 12/20-mesh. An alternate approach using
a 20/40-mesh proppant stage ahead of the 12/20 may be considered.
Graphical Output
Fig. 25, Fig. 26, Fig. 27, Fig. 28, and Fig. 29 provide the basis of the graphical output
changes for the PLACEMENT II simulator. Most of the figures are self-explanatory.
Fig. 25, Stage Front Propagation, shows the fluid front development versus time.
For this example, the 20/40-mesh proppant bridged at approximately 100 ft (7.4 min
into the treatment). The pad and only the fluid portion of the 2 PPA and 4 PPA fronts
(less proppant) continued to move through the pack during the remaining pump time
until the leakoff rate in front of the proppant pack exceeded the fluid rate through the
pack. The fluid front then receded toward and across the pack until pumping
ceased.
Note —When a screenout occurs near the fracture tip, fracture length and leakoff
area become restricted. Fracture fluid efficiency increases due to the elimination of
spurt losses and the existence of a well-developed filter cake. The use of a higher
leakoff coefficient for the slurry (as compared to the pad fluid) should be considered.
This will allow rapid concentration of the proppant and will provide reduced fracture
closure time with minimum proppant settling. PLACEMENT II simulation time may
be extended significantly if closure time is long.

3 Execution
The differences between high-permeability fracturing and conventional fracturing are
reduced treatment sizes, batch mixing requirements and lack of conventional
fracturing equipment in a “non-frac” area. Short pump times will enable the use of
limited-stability fluids using the CleanFRAC Service with high breaker concentrations
to maximize conductivity. Larger treatments designed for tip screenout will require
continuous-mix POD* blender capability. Logistics in many offshore environments
may dictate equipment constraints.

3.1 Batch-Mix Operations


Batch-mix operations are applicable for the following.
• Skin-bypass treatments when proppant slurry volumes are generally less than
4000 gal
Offshore environments and remote areas with limited access to fracturing
equipment can use existing pumping and batch-mixing capability. As few as

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 55 of 76

one 250 HHP pump has been used to successfully fracture at 5 bbl/min
placing 2000 lbm of proppant at 8.5 PPA in YF140.
• Pump rates limited to less than 12 bbl/min
At lower pump rates (less than 4 bbl/min per pump), recirculation of batch-
mixed slurry through suction manifolding may be required for linear fluids. For
batch-mix treatments at higher rates (greater than 12 bbl/min) modifications to
slurry batch-mix centrifugal pumps and manifold equipment (six to eight in.
diameter) may be required.
• When the concentration of proppant is generally limited to one (8 to 12 PPA) or
two (2 to 4 and 8 to 12 PPA) stages.
Tip-screenout design is more difficult to apply due to lack of volume and
ramping capability.

3.2 Continuous-Mix Operations


Conventional fracturing equipment has advantages in the execution of the tip-
screenout design.
• The POD blender should be used for accurate control of the short- and high-
concentration proppant ramps.
• A stage of low-concentration proppant (1 to 2 PPA) ahead of the ramp to
establish fracture entry and initiate the tip screenout is recommended as part of
the design. A low-concentration stage ahead helps to minimize “pack-back” of
the proppant. This stage also allows time to switch the blender to automatic
control.
• When confidence in the input parameters or design model is low, a provision may
be made to lengthen, at fixed concentration, the stage during which the tip
screenout is predicted. If net pressure is calculated from surface pressure, the
stage should be sized to fill the wellbore and tubulars so that screenout pressure
rather than friction pressure is reflected.
• In the event fracture-tip screenout and a net pressure increase is not evident
during the remaining 10% to 20% of slurry volume and displacement, rate
reductions may be used to reduce hydraulic width, accelerate fracture closure
and avoid fracture propagation after shut down. Rate reductions during a net
pressure increase may also be required to avoid overpressure.

4 Evaluation
In moderate- to high-permeability reservoirs, evaluation of finite-conductivity
fractures may not be possible using conventional type-curve, and bilinear-flow
analyses of post-fracture pressure-buildup data. The existence of the characteristic
one-quarter slope on the log-log diagnostic plot of differential pressure and derivative
versus time is required, and its presence provides an indication of stimulation.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 56 of 76

Gas well evaluation should consider the effects of non-darcy flow during the buildup
period and the resulting apparent fracture conductivity variance with time
(Holditch/Morse, Tannich/Nierode). A numerical simulator accounting for non-darcy
flow effects must be used to history-match the data for these conditions.

4.1 Prats’ Correlation


1. In the majority of high-permeability cases, the short bilinear flow period will not
be observed and a pseudo-radial flow regime may be analyzed (Horner plot) to
determine the effective skin.
2. The effective wellbore radius may then be estimated using Eq. 4 and compared
to predicted value.
3. The predicted apparent dimensionless fracture conductivity and are used to
obtain the dimensionless ratio of rw′/xf.
4. With the calculated effective wellbore radius and the ratio known, the effective
fracture length may be computed.

4.2 Modified McGuire-Sikora Correlation


1. An alternate approach using Fig. 4 may be used. The ratio of undamaged
productivity potential (Eq. 16 and Eq. 17) versus the actual observed rates,
defines the pseudo-steady-state folds of increase.
2. From the appropriate well spacing and folds increase value, a horizontal line
may be drawn to the intersection of the predicted dimensionless fracture
conductivity (or conductivity ratio kfw /kre) and corresponding xf /re.
3. The fracture length may then be computed from the xf /re value and compared to
the predicted length.
Evaluation using pseudo-steady-state flow rates will be sensitive to the average
differential pressure used to compute rate. Factors such as depletion (avg. Pe) and
non-darcy pressure losses should be considered (numerical reservoir simulation and
history matching). In some cases, limited drawdown may introduce inaccuracies in
differential pressure for rates quoted.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 57 of 76

Fig. 17. PLACEMENT SIMULATOR — conventional design,


20/40-mesh sand, 1400 gal pad.

Fig. 18. PLACEMENT OUTPUT — conventional design, 20/40-mesh sand.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 58 of 76

Fig. 19. PLACEMENT SIMULATOR — P3D tip-screenout design,


20/40-mesh sand, 1600 gal pad.

Fig. 20. PLACEMENT OUTPUT — P3D tip-screenout design, 20/40-mesh sand.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 59 of 76

Fig. 21. PLACEMENT SIMULATOR — conventional design,


12/20-mesh sand, 1800 gal pad.

Fig. 22. PLACEMENT OUTPUT — conventional design, 12/20- mesh sand.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 60 of 76

Fig. 23. PLACEMENT SIMULATOR — P3D tip-screenout design,


12/20-mesh sand, 3500 gal pad.

Fig. 24. PLACEMENT OUTPUT — P3D tip-screenout design, 12/20-mesh sand.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 61 of 76

Fig. 25. Stage front propogation.

Fig. 26. Fracture height profile

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 62 of 76

Fig. 27. Wellbore fracture width profile.

Fig. 28. Fracture height growth history.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 63 of 76

Fig. 29. Fracturing (net) pressure profile.

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 64 of 76

5 Fluid-Loss Data

5.1 WF120 (J164) Containing 25 lbm J478/1000 gal and 25 lbm J418/1000 gal
BHST=150°°F (66°°C), Pressure=1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 65 of 76

5.2 WF160 (J164) Without Fluid-Loss Additives  BHST=150°°F (66°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 66 of 76

5.3 WF160 (J164) Containing Various Fluid-Loss Additives  BHST=150°°F (66°°C),


Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 67 of 76

5.4 WF110 (J424) Containing 50lbm J238/1000 gal  BHST=150°°F (66°°C), Pressure-
1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 68 of 76

5.5 WF120 (J424) Containing Various Fluid-Loss Additives  BHST=150°°F (66°°C),


Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 69 of 76

5.6 WF130 (J424) Containing 50lbm J238/1000 gal  BHST=150°°F (66°°C),


Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 70 of 76

5.7 WF140 (J424) Containing 50lbm J238/1000 gal  BHST=150°°F (66°°C),


Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 71 of 76

5.8 WF160 (J424) Containing Various Fluid-Loss Additives  BHST=150°°F (66°°C),


Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 72 of 76

5.9 YF140 (J424)  BHST=150°°F (66°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 73 of 76

5.10 YF140 (J424)  BHST=150°°F (66°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 74 of 76

5.11 YF140 (J424)  BHST=175°°F (79°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
HyPerSTIM Service
Page 75 of 76

5.12 YF140HTD (J424)  BHST=200°°F (93°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 800
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
HyPerSTIM Service Dowell
Page 76 of 76

5.13 YF140HTD (J424)  BHST=250°°F (121°°C), Pressure-1000 psi

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 1 of 37

ACID FRACTURING

1 Principles of Acid Fracturing.................................................................................................. 2


1.1 Fracture Length and Fracture Conductivity.......................................................................... 4
1.2 Factors Affecting Acid Behavior in Carbonate Reservoirs ................................................... 6
1.2.1 Acid Type, Strength, and Volume ............................................................................... 6
1.2.2 Acid Leakoff ................................................................................................................ 9
1.2.3 Controlling Acid Leakoff............................................................................................ 14
1.2.4 Acid Reaction Rate ................................................................................................... 15
1.2.5 Acid Spending Time.................................................................................................. 17

2 Treatment Design Fundamentals for Acid Fracturing........................................................ 18


2.1 Achieving Acid Penetration ................................................................................................ 19
2.2 When Acid Fracture Length Should be Maximum.............................................................. 19
2.3 When Acid Fracture Length Should be Limited.................................................................. 19
2.4 Maximizing The Injection Rate ........................................................................................... 20
2.5 Optimizing Conductivity and Etched Fracture Length ........................................................ 20
2.6 Effective Acid Concentration .............................................................................................. 21
2.7 Selecting Fluids for Deeper Acid Penetration .................................................................... 22
2.8 Determination of Leakoff Coefficients ................................................................................ 24
2.8.1 Methodology ............................................................................................................. 24
2.8.2 Example Calculation ................................................................................................. 29
2.8.3 Notes ........................................................................................................................ 31
2.9 Cooldown ........................................................................................................................... 31
2.10 Retarded Acid .................................................................................................................. 33
2.11 Viscous Fingering ............................................................................................................ 36
2.12 Summary of Treatment Design Fundamentals for Acid Fracturing .................................. 37

FIGURES

Fig. 1. Conductivity ratio versus increase in folds. ....................................................................... 5


Fig. 2. Acid spending in carbonate rock. ...................................................................................... 8
Fig. 3. Casting of a typical wormhole pattern in limestone......................................................... 12

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 2 of 37

Fig. 4. Test results, carbonate cores and wormholes.................................................................12


Fig. 5. Test results, carbonate cores and wormholes.................................................................13
Fig. 6. Test results, carbonate cores and wormholes.................................................................13
Fig. 7. Viscosity of fresh water versus temperature....................................................................25
Fig. 8. Spending time versus HCl concentration at static conditions. .........................................35
Fig. 9. Temperature versus retardation factor at dynamic test conditions..................................36

TABLES

Table 1. Factors that Influence Fracture Conductivity and Fracture Penetration when Acid
Fracturing a Carbonate Reservoir ..................................................................................6
Table 2. Acid Types and Strengths Common to Oilfield Operations ............................................9
Table 3. Total Minimum Leakoff Coefficients versus Temperature and Permeability ................26
Table 4. Scale Factor Base Values (SFb) ...................................................................................27
Table 5. Scale Factor Corrections ..............................................................................................28
Table 6. Retardation Factor Selection Guidelines ......................................................................34

1 Principles of Acid Fracturing


Acid fracturing, also called fracture acidizing, is a stimulation process in which acid,
usually hydrochloric acid (HCl), is injected into a limestone or dolomite formation at a
pressure sufficient to fracture the formation or to open existing fractures. As the acid
flows along the fracture, portions of the fracture face are dissolved. Since flowing
acid tends to etch in a non-uniform manner, conductive channels are created which
usually remain after the fracture closes. The length of the etched fracture is
determined by the acid type, strength, volume, acid leakoff parameters, reaction rate
and spending rate. These factors are mutually dependent upon each other.
The effectiveness of the acid fracturing treatment is largely determined by the length
of the etched fracture. As with conventional hydraulic fracturing operations, the
principal objective of an acid fracturing treatment is to provide a conductive fracture
with sufficient length to allow efficient drainage of the reservoir. The major difference
between the two treatment techniques is how the conductivity is achieved. In
conventional propped fracturing treatments, sand or other proppant is placed in the
fracture to prevent closure when fracturing pressure is released. Acid fracturing
does not normally utilize propping agents but relies upon acid-etched fracture faces
to provide the necessary conductivity.
Acid fracturing is generally limited to the cleaner, higher solubility, limestone or
dolomite (carbonate) formations. Dirty carbonate rocks (less than 70% solubility in
HCl) are not candidates for acid fracturing for the following reasons:
• The creation of acid-etched flow channels will be impaired because of low
solubility.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 3 of 37

• The release of insoluble materials will tend to plug any conductive etch patterns
created by the acid.
Although normally possessing high acid solubility, chalk formations may not be
suitable acid fracturing candidates because of the soft character of the rock. The
rock must be sufficiently competent to retain conductivity after closure of the etched
fracture.
Acid fracturing is never used in the treatment of sandstones because acid, even
hydrofluoric acid (HF), will not adequately etch sandstone fracture faces. Even if the
sandstone is cemented with a carbonate cement, materials released through the
dissolution of the carbonate cement will plug the fracture.
In some carbonate formations, a choice exists between acid fracturing and proppant
fracturing. Each technique has advantages and disadvantages. If similar
productivity can be achieved by either procedure, the choice is usually a nonreactive
fracturing fluid with proppant because of economic considerations. Acid fracturing of
reservoirs having bottomhole static temperatures in excess of 250°F (121°C) can be
expensive because of increased corrosion inhibitor requirements, especially on
steels normally used under such conditions (13% chrome, duplex). A tradeoff of
costs can be achieved by using the cooldown effect of the pad on the well tubulars
and adding only sufficient corrosion inhibitor to provide corrosion protection at the
cooldown temperature. The savings in inhibitor costs may be negated by the costs
of additional and appropriate standby equipment to continue and complete the job,
should it become necessary.
Operationally, acid fracturing is less complicated because propping agent is not
used. Also, the danger of screenout and the associated problems of proppant
flowback and cleanout from the wellbore are eliminated. In addition, the fracturing
fluid proppant transport characteristics are not a concern with acid fracturing. In
deep carbonate wells, acid may be the best fluid for fracturing operations if proppant
flowback, crushing or screenout are a problem, despite the corrosion problems
mentioned above.
While the use of acid as a fracturing fluid eliminates some difficulties inherent with
proppant fracturing, there are problems of a different nature. For example, when
comparing a propped fracture and an acid-etched fracture, the effective length of the
propped fracture is limited by the distance the proppant can be transported down the
fracture. In a similar manner, the effective length of the acid-etched fracture is
limited by the distance the acid can travel along the fracture and adequately etch the
fracture faces before becoming spent. The effective acid penetration is always
shorter than any proppant placement, especially at elevated temperatures.
An acid fracturing design should take into consideration the following:
• The fracture should be propagated to the desired length in the reservoir.
• The walls of the fracture should be etched with an acid capable of dissolving
large amounts of reservoir rock.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 4 of 37

• The etched fracture should retain adequate length and conductivity after closure.
• Rapid cleanup of treatment fluids should be achieved.
• The acid fracturing treatment should be the most cost-effective (best value)
treatment for the client.
Summary
The best application for acid fracturing may be in low-permeability carbonate
reservoirs having a limited number of natural fissures. Chalk formations can be a
problem because of soft characteristics (the acid-etched fracture may collapse).
Proppant fracturing may be best in deep, carbonate reservoirs because the depth,
high temperatures, and high closure pressures are detrimental to the length and
conductivity of an acid-etched fracture. In addition, corrosion inhibition costs are
usually high in these situations.

1.1 Fracture Length and Fracture Conductivity


The objective of an acid fracturing treatment, like a conventional proppant fracturing
treatment, is to create a conductive fracture with sufficient length to provide effective
drainage of the reservoir.
The McGuire and Sikora stimulation curves (Fig. 1) relate the fundamentals of
relative conductivity and fracture penetration to the drainage radius. The relative
conductivity is a comparison of the induced fracture permeability and the
unstimulated reservoir permeability. The curves indicate that a high conductivity
ratio and a deeply penetrating flow channel are both necessary for significant
productivity increase. Alone, neither of them will provide stimulation. These
principles (conductivity and penetration) apply whether proppant fracturing or acid
fracturing.
When acid fracturing, the etched length, not the hydraulic length, is considered the
fracture length.
Fracture conductivity, when acid fracturing, is created by dissolving reservoir rock.
When a given volume of acid is pumped into a carbonate reservoir, a specific
amount of reservoir rock will be dissolved by the acid. The created void represents
conductivity. In order for an acid fracturing technique to be successful however, the
conductivity must be optimized by assuring that reservoir rock will be dissolved in a
manner that will provide highly conductive flow channels.
Hydrochloric acid reacts with carbonate rock in a nonuniform manner, that is, finger-
like channels are created along the fracture faces. The area between these
channels serve as pillars. When the fracture closes, the channels remain, providing
high conductivity.
Acid fracture conductivity and acid fracture penetration (both are necessary for
reservoir stimulation - see Fig. 1) are governed by many factors. An understanding

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 5 of 37

of these factors and how they interrelate is necessary when considering stimulation
by acid fracturing. These factors are shown in Table 1.
There are five prerequisites for reservoir stimulation using acid.
1. The acid must dissolve reservoir rock and form soluble by-products that can be
returned from the well.
2. Corrosion inhibition must be relatively inexpensive.
3. Components of the acid fracturing treatment must be relatively safe to handle.
4. The acid fracturing fluid must be readily available.
The acid fracturing treatment must be relatively inexpensive.

Fig. 1. Conductivity ratio versus increase in folds.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 6 of 37

Table 1. Factors that Influence Fracture Conductivity and Fracture


Penetration when Acid Fracturing a Carbonate Reservoir
• Acid Type
• Acid Strength
• Acid Volume
• Acid Leakoff
• Acid Reaction Rate (Acid Spending Time)
• Acid Viscosity
• Injection Rate
• Formation Mineralogy
• Formation Temperature
• Formation Wettability
• Reservoir Fluids
• Fracture Net Pressure
• Closure Pressure

1.2 Factors Affecting Acid Behavior in Carbonate Reservoirs

1.2.1 Acid Type, Strength, and Volume


The most common acid used in the oilfield for acid fracturing is hydrochloric acid.
HCl is a strong mineral acid.
Weak organic acids such as formic (HCOOH) and acetic (CH 3COOH), are also used.
The dissolving power of organic acids is much lower than that of HCl. Organic acids
do, however, have the characteristic of being easier to inhibit against corrosion than
HCl. Organic acids have been used in blends of hydrochloric and formic acids or
hydrochloric and acetic acids. These blends have been thought to provide a degree
of retardation of the acid so that live acid could be pushed further down the fracture,
creating conductivity deeper in the reservoir. Under most conditions, however, when
the HCl spends, the weak organic acid, although unspent, lacks the reactivity to
adequately etch the fracture faces.
Formic acid, acetic acid and organic acid mixtures have been used to fracture
reservoirs with bottomhole static temperatures greater than 300°F (149°C).
Development of more efficient corrosion inhibitors along with well cooldown
procedures have diminished the use of formic and acetic acids. They are however
sometimes used as the breakdown fluid in high temperature reservoirs prior to
fracturing operations using proppant. Corrosion inhibition of organic acids at high
temperatures is easier than corrosion inhibition of HCl.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 7 of 37

Chemically, whether an acid is defined as a strong acid like HCl or as a weak acid
like either formic or acetic depends upon the rate at which the acid ionizes as an
electrolyte. An electrolyte is an aqueous solution that has the capability of
transmitting an electric current. The defining parameter to establish whether an acid
is strong or weak is the rate at which its acid solution transmits electric current.
Consider a hypothetical acid, HA, which has an acid ion H+ and a negative ion A-.
Whether it is a strong acid or a weak acid depends upon how fast its acid ion, H+,
dissociates (ionizes) from the negative A- ion in a water solution.
HCl, in water, instantly and completely ionizes into its component ions (H+,Cl-).
Instant ionization takes place regardless of the acid concentration in water from 3 to
37% (by weight). Thus, to be technically correct, lesser concentrations of HCl (for
example, 3% to 5%) are more properly defined as dilute acid solutions rather than
weak acid solutions. Because total dissociation of HCl is instant, its ionization
constant is infinite. Formic acid, in water, ionizes into its component ions slower than
HCl but faster and more completely than acetic acid. Both are much weaker acids
than HCl. Being defined as weak acids does not mean they are less hazardous to
health, instead, `weak' defines the degree of ionization. Both these organic acids
are very corrosive to flesh. Material Safety Data Sheet provide safe handling
procedures for the various acids.
The chemical equation for reaction of hydrochloric acid with limestone (CaCO3) is
shown below in mass-balanced form. Note that the reaction products are soluble in
water and can be readily recovered from the well.
CaCO3 + 2 HCl → H2O + CO2 ↑ + CaCl2
(Limestone) + (Acid) → (Water) + (Gas) + (Salt)
The reaction products are water, carbon dioxide gas and calcium chloride salt.
Hydrochloric acid is the most common type of acid used in the oil field. HCl is
variously inhibited, stabilized, retarded, gelled, blended with other acids, foamed,
misted, emulsified and nonemulsified. Other additives include friction reducers,
surface-tension reducers, mutual solvents, chelants, fluid-loss control additives,
water-block removers, viscosifiers, scale inhibitors, paraffin inhibitors, additives to
cause heating, diverting agents, clay stabilizers and clay dispersers. The usual HCl
concentrations in acid fracturing are 15, 20 and 28%.
The other two familiar types of acids used in the oil field, formic acid and acetic acid,
also react with limestone to form water, gas and a salt. These weak organic acids
do not completely spend in carbonate reservoirs but buffer themselves with the
reaction products present in the partially spent acid. Depending upon reservoir
temperature and acid concentration, these organic acids will spend to a certain point
and no further. (See Fig. 2 and Table 2. HCl is shown only for reference. It spends
to completion at all temperatures.) The effect of higher temperatures on organic
acids is to decrease reactivity in carbonate reservoirs. Additionally, neither of these
organic acids should be used at concentrations above 9% formic or 10% acetic

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 8 of 37

since, when spent, high concentrations of organic calcium salts may precipitate in
the formation:
Formic Acid
CaCO3 + 2 HCOOH → H2O + CO2↑
+ Ca(HCOO)2 (Calcium formate)
Acetic Acid
CaCO3 + 2 CH3COOH → H2O + CO2↑
+ Ca(CH3COO)2 (Calcium acetate)

Fig. 2. Acid spending in carbonate rock.


Table 2 lists several different acid types and strengths common to oilfield operations.
Note that a considerable amount of CO2 can be generated by the acids. Gas
expansion is a potential energy that can be used to help recover the treatment load
by back-flowing the well as soon as possible after shutdown of the fracture acid
treatment.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 9 of 37

As stated previously, a given amount of acid will create a given amount of etched
fracture width. Higher volumes and higher strengths will create even more etched
fracture width. This can be translated into useful fracture conductivity providing that
the rock that has been dissolved is in the form of deeply penetrating flow channels
etched into the fracture faces.

Table 2. Acid Types and Strengths Common to Oilfield Operations


Acid Type Acid lbm CaCo3 Dissolved scf CO2
Strength (%) by 1000 gal Acid Generated per
bbl Acid
Hydrochloric 15 1833 289
20 2515 396
28 3662 577
Formic+ 9 726 60
Acetic+ 10 422 30
+at 100° F(38°C)

1.2.2 Acid Leakoff


During many acid fracturing treatments, the treating pressure declines continually,
eventually falling below a level required to propagate the fracture. This decline in
treating pressure is caused by the usually very large leakoff of the acid. The overall
leakoff rate increases with fracture length, up to a point where it equals the injection
rate and where net pressure is minimum and the fracture stops extending.
Conventional hydraulic fracturing using nonreactive fluid relies primarily on wall-
building fluid-loss additives to control leakoff. The filter cake deposited on the face of
the fracture by the fluid-loss additive allows less and less fluid to leak off into the
formation. This same kind of control is desirable when acid fracturing, but, because
conventional filter cakes are destroyed by the acid or are negated by the reaction of
acid dissolving rock from beneath the filter cake, other means have been developed
to combat acid leakoff.
Natural Fissures and Fractures
Carbonate reservoirs, especially dolomites, are normally more naturally fissured that
sandstones. Fissures represent a particular path of least resistance for acid. The
fissures get wider as more acid is introduced and thief away large volumes of acid
from the induced hydraulic fracture. A reservoir may contain so many fissures that
acid leakoff into them can limit the propagation of a hydraulic fracture to no more
than tens of feet. In these reservoirs, a hydraulic fracture may be initiated but can
not grow to extend past the first intersecting fissures.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 10 of 37

When the permeability of the fissures has the same order of magnitude as the matrix
permeability, the reservoir has a double-porosity behavior. Pressure tests show two
distinct periods of time, the earlier one related to the network of fissures, the second
to the matrix itself. The overall reservoir permeability is then generally equal to 2 to
20 times the matrix permeability (the latter being determined through core testing).
When dual-porosity behavior is suspected, plan first for controlling leakoff into the
fissures. Both the viscosity-controlled leakoff and the compressibility-controlled
leakoff are proportional to the square root of permeability. Thus if matrix
permeability is 0.1 md and reservoir permeability is 1.7 md, control of leakoff into
fissures reduces the leakoff rate by a factor of four.
When fissure permeability overwhelms the matrix permeability, the reservoir tends to
behave again as a primary porosity medium. In this latter case, it would be difficult
to initiate a hydraulic fracture and any such fracture would provide very little
production increase. However, high-rate acid treatments are common in these
reservoirs. They make use of Mud and Silt Remover (MSR*)-type acids, NARS*
formation solvent or highly concentrated surfactant pills, and are aimed at removing
drilling mud losses from the fissures network. Such treatments are not acid
fracturing treatments and cannot be designed using the present guidelines.
Wormholes
During an acid fracturing treatment, the acid not only etches the fracture faces, it
also leaks off into the reservoir perpendicular to the fracture faces. Live acid leaking
into the porous fracture face creates voids called wormholes. These become
irregular, meandering, highly conductive tubes that can penetrate several feet into
the reservoir rock (see Fig. 3). They are caused by selective enlargement of the
larger pores in the rock as it reacts with the acid. Once wormholes develop, more
acid leakoff occurs primarily via these wormholes. There is little leakoff in the
conventional sense. Wormholes divert large volumes of acid away from the primary
fracture system, volumes which are then unavailable to etch the fracture face a
further distance from the wellbore. The magnitude of leakoff due to the wormhole
effect is very severe. Wormholes occur in porous limestone and dolomite
formations.
The same factors that govern the acid penetration along the hydraulic fracture also
affect acid penetration into pores. Consequently, wormhole enlargement and etched
fracture length are affected by many of the same factors. For instance, fracture
etched length increases as injection rate and fracture width increase. Similarly,
wormholing is enhanced by increased leakoff rate and increased pore diameter.
(Note however, that when matrix acidizing, wormhole diameter may actually
decrease as injection rate increases. This is because of the large number of
wormholes that are attempting propagation at the same time and the extremely large
area of rock that is contacted by a limited acid volume).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 11 of 37

This helps explain why wormholes form in the porosity of the rock matrix. The
combination of large pore diameters and the high rate of acid molecules within the
acid flow inhibits the reaction of acid at the pore inlet and, because the acid molecule
tends to stay in the center of the flowing stream, supports the extension of deeper
wormholes. This also helps explain why, as temperature increases, wormholes are
less of a problem. The acid reaction rate is increased to the point that the acid will
spend at the inlet of the rock pores, whatever the pore size. This occurs at
approximately 225°F (105°C) in limestones and approximately 325°F (163°C) in
dolomites when pore diameters are in the 0.5 to 2 micron range which is normally
1 to 10 md permeability. For lower matrix permeabilities, and therefore smaller pore
diameters, the temperature limits shift considerably, down to approximately 100 °F
(38°C) in limestones below 0.1 md.
A compact carbonate reservoir with very low matrix permeability (below 0.1 md)
does not experience wormholing under normal treating pressures, that is, normal
differential pressure between fracture and reservoir. Conversely, hairline fissures
with widths of several microns are as many initiators for wormholing at elevated
temperatures, even up to 350°F (177°C).
Fig. 4, Fig. 5 and Fig. 6 illustrate the results of laboratory testing of carbonate cores
and wormholes created with HCl. The separate effects of temperature, rate and
concentration are plotted versus wormhole penetration of similar Indiana limestone
cores. For these tests, the diameter of the cores was one in. and length was 11.8 in.
To summarize, the main effect of acid leakoff is to reduce the etched fracture length
and provoke a significant reduction in real versus planned well performance. This is
true in spite of the increased wormhole-induced permeability of reservoir rock along
the induced hydraulic fracture near the wellbore.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 12 of 37

Fig. 3. Casting of a typical wormhole pattern in limestone.

Fig. 4. Test results, carbonate cores and wormholes.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 13 of 37

Fig. 5. Test results, carbonate cores and wormholes.

Fig. 6. Test results, carbonate cores and wormholes.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 14 of 37

1.2.3 Controlling Acid Leakoff


Fluid-loss control during acid fracturing of carbonate formations presents problems
unique to reactive fluids. Most fluid-loss additives and gelling agents that are
effective in conventional nonreactive aqueous fracturing fluids are unstable in acid.
As a result, special acid-stable additives must be used. In addition, acid flow along
carbonate fracture surfaces produces constant chemical erosion, thus making it
difficult for wall-building fluids to deposit an effective filter cake. Finally, the acid
leakoff rate continuously increases all along the acid injection due to the
development of wormholes.
Various additives and treating techniques have been developed as means of
controlling acid fluid loss. These include acid-swellable polymers, oil-soluble resins,
injection of gelled water pad ahead of the acid, injection of gelled water pads
alternately spaced within stages of acid and acid-gelling agents.
Natural and Synthetic Swellable Polymers
Natural and synthetic swellable polymers have been used with limited success.
These polymers do not readily dissolve in acid but form small swollen particles that
act to block wormholes in the early stage of development. Natural swellable
polymers degrade rapidly in acid at temperatures greater than 125°F (52°C).
Synthetic, swellable polymers, developed to overcome this problem, perform well in
the laboratory but field treatments sometimes result in poor or incomplete post-
treatment cleanup.
Oil-Soluble Resins
To be effective, additives must be stable in acid solutions at elevated temperature.
Fluid-loss additives that are stable in hot acid, however, are generally very difficult to
dissolve or degrade after the treatment. One method of overcoming this problem
involves the use of mixtures of oil-soluble resins as acid fluid-loss additives. The
major limitation of oil-soluble resins is the high concentration of additive required for
fluid-loss control. At the required concentration of 200 lbm resin mixture/1000 gal
acid, high additive-costs limit commercial application.
Injection of a Viscous, Nonreactive Fluid Ahead of the Acid
The "Frac Pad and Acid" technique involves using a water-base fracturing fluid to
initiate and cool the fracture and deposit an impermeable filter cake on the fracture
face. Though widely used, tests have indicated that the filter cake is quickly eroded
and penetrated by acid. Once this occurs, acid fluid loss is identical to that observed
if no gelled pad were used.
Pads of linear gels that do not build polymer cakes on fracture face produce a
viscous filtrate which might be very efficient at controlling leakoff during the following
acid stage. This is particularly true in gas reservoirs (even more when depleted)
where the viscosity, compressibility coefficient (Cvc) is naturally very large.
While the injection of a viscous, nonreactive fluid ahead of the acid provides only
very limited acid fluid-loss control, there are other benefits. The tubulars through

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 15 of 37

which the acid must flow are cooled, reducing corrosion. The technique cools the
fracture and increases fracture width, reducing acid spending rate and increasing
live-acid penetration along the fracture. Additionally, using a viscous pad promotes
acid fingering of the acid into the trailing edge of the pad, thus reducing the amount
of reactive surface to which the acid is exposed, improving etched fracture length.
Injection of Gelled-Water Pads Alternately Spaced With Stages of Acid
This procedure was developed by Dowell and is marketed as the DUOFRAC* II acid
fracturing procedure. Using this procedure, the fracture is created by a gelled pad,
after which alternating stages of acid and additional pad are pumped. The additional
pad stages are designed to enter and seal wormholes created by the preceding
stage of acid. Acid leakoff into wormholes is slowed and treatment efficiency is
improved. The DUOFRAC II acid fracturing procedure has all the advantages of the
Frac Pad and Acid technique and, in addition, controls acid leakoff to a greater
degree.
A fine particulate solid is often added to the pad stages to aid in fluid-loss control.
Examples are 100-mesh sand, oil-soluble resins, and fine salt. These particulate
solids help bridge wormholes and natural fractures, also reducing leakoff.
Gelled Acid
Acid leakoff can also be reduced by gelling the acid. A viscous fluid has less
tendency to invade pore openings than a water-thin fluid. This method of control has
become widely used with the development of more acid-stable thickening agents.
The effectiveness of these thickeners varies widely. Some viscous acid systems,
like Leakoff Control Acids (LCA* fluids), are very effective.
LCA fluids are excellent acid fracturing fluids. These polymer-base gelled acids
have low initial viscosity and friction pressure properties. During leakoff, LCA fluids
temporarily develop high viscosity which controls fluid loss by blocking wormhole
growth and slowing acid entry into natural fractures. LCA fluids have low spent-acid
viscosity which enhances cleanup.
The Fracturing Materials ManualFluids provides additional information on high-
performance leakoff-control fluids.
Gelling agents such as guar gum, hydroxypropylguar (HPG), hydroxyethylcellulose
(HEC), and carboxymethylhydroxyethylcellulose (CMHEC) have limited stability in
HCl and are thus limited in application for gelled acid fracturing.

1.2.4 Acid Reaction Rate


The definition of acid reaction rate is the number of acid molecules reacting with the
carbonate rock per unit of time. Acid reaction rate is not to be confused with acid
spending time. Reaction rate is the speed at which acid reacts. Spending time is
the length of time required for the acid to spend.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 16 of 37

There are two different mechanisms involved in the overall reaction rate of acid.
1. The physical transportation of acid molecules (diffusion).
2. The chemical reaction of acid and rock molecules (kinetic).
Both mechanisms can be slowed down by decreasing the temperature of the acid
mixture; surface reaction by adsorbing protective layers on the rock surface.
Diffusion
Physical transport of the acid molecules from the center of the induced fracture
toward the fracture face involves diffusion-rate and convection-rate phenomena.
Together, these two terms are called the mass-transport rate.
Diffusion is the natural motion of the molecules of any fluid that depends upon
temperature and their spontaneous movement from a region of higher to a region of
lower concentration.
Convection is the motion of molecules that occurs in a fluid submitted to gradients of
temperature or pressure (thermic and hydraulic convections respectively).
• Molecular diffusion is slowed by increasing the interaction between molecules
and by putting obstacles in the path of the molecules. This may be achieved by
the physical presence of products generated by the acid reaction. (See Kinetic
on the following page, which presents a discussion of the chemical reaction of
acid and carbonate rock.) Diffusion can also be reduced by decreasing the
temperature.
• Turbulence enhances random motion of molecules, thus if turbulence is
suppressed, diffusion is slowed. Turbulence can be reduced by acid viscosifiers,
friction reducers and wider fractures.
• Reducing the acid leakoff rate reduces convection by slowing the movement of
the acid molecules toward the fracture faces.
• As the fracture width is increased, more time is required for a molecule having a
given velocity to reach the fracture face.
• Reducing the surface-reaction rate can be accomplished either by reducing the
temperature or by insertion of a barrier between the acid and the rock.
Temperature reduction can be achieved using cool-down techniques. A barrier
can be in the form of an acid-resistant filter cake, presence of reservoir oil or any
kind of hydrocarbon barrier material used in the treating fluid. Examples of
relevant Dowell treatment fluids are Acid Retarder F98 retarded acid (F98 is a
surfactant which strongly adsorbs on carbonate surfaces) and SuperX* Emulsion
(the external oil or diesel phase acts as a barrier in the bulk of the fluid itself).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 17 of 37

Kinetic
The chemical reaction of the acid molecules at the fracture face results in the
recombination of the hydrochloric acid and the rock species into water-soluble salts,
carbon dioxide, and water. When HCl acid reacts with dolomite, the recombination
products are calcium and magnesium chlorides (CaCl2 and MgCl2), carbon dioxide,
and water. The presence of these compounds in solution in the partially spent acid
will help suppress movement of the acid molecules toward the fracture face.
The chemical reaction rate is a function of the interaction of all factors previously
discussed including temperature.

1.2.5 Acid Spending Time


Acid spending time is defined as the time taken for acid to spend under treating
conditions. In contrast, the term acid reaction rate (previously discussed) defines the
rate that acid spends. In many cases, it may not be necessary to retard the reaction
rate of HCl but simply to cause it to spend at the desired locations within the fracture.
In other words, increase the effectiveness of the acid by causing it to spend within
the fracture.
The acid fracturing concept is basic: fracture with acid to dissolve rock along the
fracture faces resulting in greater conductivity. Placing the acid in the right place
along the induced fracture so the desired flow channels will be created is important.
The time required for HCl to spend during acid fracturing is dependent upon many
parameters. Since these parameters are constantly changing because of the
chemical erosion of carbonate rock, these parameters affect each other. Each of
these parameters; strength and volume of acid, acid leakoff rate, live and spent acid
viscosities, acid injection rate, area-to-volume ratio, temperature and differential
pressure contribute to the reaction. The reservoir properties of mineralogy, lithology,
saturation and wettability also affect spending time.
Area-to-volume ratio is defined as the ratio of surface area of rock in contact with a
specific volume of acid. A large area-to-volume ratio shortens acid spending time;
conversely, a small area-to-volume ratio will extend acid spending time.
Formation mineralogy and lithology have an effect on spending time as do reservoir
saturation and reservoir wettability.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 18 of 37

Mineralogy
Different acid-soluble minerals exhibit different spending times even if other
conditions are the same. Calcite and dolomite, each having a different spending
time, are the predominant minerals found in carbonate formations and are the
minerals that are targeted for dissolution. Other acid-soluble minerals include other
carbonates, iron sulfides, halide salts (soluble in the acid make-up water), and, to a
limited extent, clays and feldspars.
Lithology
Lithology defines the physical properties of the reservoir rock, that is, the degree of
mineral crystallization within the reservoir, the sequence of deposition, bedding
within the rock and presence of other rock particles.
Oil-Saturated Reservoirs
An oil-saturated reservoir will react more slowly with the acid than a water-saturated
reservoir because the oil can form a temporary barrier between the acid and the
rock.
Oil-Wet Reservoirs
An oil-wet reservoir can have a greater barrier-effect than one that contains high oil-
saturation.
None of these reservoir factors are readily or significantly changeable. Thus, the
acid fracturing technique should be tailored in a way that can reduce or negate
detrimental reservoir parameters.

2 Treatment Design Fundamentals for Acid Fracturing


This section contains information that can be of assistance to users of the
FracCADE* software for the design of an acid fracturing treatment. This information
is provided so the reader can have a better understanding about acid fracturing and
its design parameters. These guidelines are not intended to be a substitute for
FracCADE documentation and procedures.
When determining the relative merits of different concentrations and different
mixtures of acids and treatment techniques, the FracCADE software should be used
to compare them using values derived under dynamic (flowing) conditions
representing downhole parameters. Test data derived using static conditions are of
little value. Since there are such a large number of variables (see Table 1) and an
infinite number of ways these variables can interact with each other, the acid
reaction factors given in this section should only be used to compare different acid
systems or different pumping techniques or both. Selection should be based on the
relative merits of the calculated values.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 19 of 37

2.1 Achieving Acid Penetration


The effectiveness of an acid fracturing treatment is largely determined by the length
of the etched fracture. Controlling acid fluid loss and reaction rate, with attention to
the other governing parameters such as injection rate, viscosity and area-to-volume
ratio provides the best opportunity to create a long, conductive fracture.

2.2 When Acid Fracture Length Should be Maximum


Maximum acid penetration is desired most of the time. To obtain maximum acid
penetration at a bottomhole static temperature less than 200°F (93°C), the first step
of the treatment design is to select fluids that have the lowest leakoff coefficients.
At temperatures greater than 200°F (93°C), acid reaction rate usually becomes the
primary limiting factor and must be decreased. This is achieved using reservoir
cooldown techniques or with acid retardation or both.
Higher injection rates, wider fracture widths and higher acid concentrations always
provide deeper acid penetration regardless of temperature and leakoff conditions.
Additionally, the presence of effective stress barriers bounding the producing interval
will allow lateral fracture growth. Where barriers exist, the objective is to cause the
acid-etched fracture to penetrate deeply into the drainage radius of the well.

2.3 When Acid Fracture Length Should be Limited


Carbonate reservoirs seldom contain stress barriers. Where there are no barriers to
limit fracture height growth, a hydraulically induced fracture will grow in both vertical
and lateral directions until the treatment is completed.
In the case of minimizing or preventing water or gas production from above or below
the producing interval, or in the case of preventing communication with other zones,
the fracture length should be restricted to one-half the thickness of the producing
interval, unless a fracture-height-containment technique is used. For example,
assuming a producing interval of 50 ft, the etched, conductive fracture length, that is,
the radius of the acidized fracture, should be 25 ft.
The fracture acidizing treatment must be designed to equalize the injection and
leakoff rates once the required acid penetration is obtained. This will optimize acid
etching properties. An accurate estimate of the acid leakoff coefficient is required for
that purpose and can be obtained from a pressure-decline analysis (from a previous
acid fracturing treatment in the same reservoir).
Treatments that require equalized injection and leakoff rates generally use HCl as
the sole treating fluid. To achieve the desired radial fracture extension, acid injection
rate should be above fracturing pressure until sufficient acid has been pumped.
When the desired fracture extension is reached, injection should be continued but at
a reduced rate corresponding to fracturing pressure. Continued injection at this
intermediate pressure will keep the fracture open and permit the acid to create high

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 20 of 37

conductivity over the radial length of the fracture. As pumping continues, leakoff will
increase and injection pressure will approach pore pressure.
The best results are obtained when perforations are restricted to a few holes in the
center of the producing interval.

2.4 Maximizing The Injection Rate


The first design criterion of most acid fracturing treatments is to maximize the
injection rate (jobs in cool dolomites or using LCA fluids might be exceptions). Just
as higher injection rates cause deeper penetration of the acid into the hydraulic
fracture, the faster the acid moves within the fracture, the longer it takes for a given
acid molecule to reach the fracture face and react with reservoir rock. This condition
is true as long as the acid in the fracture is moving in laminar flow.
Maximizing the acid fracturing injection rate first involves determining the maximum
allowable wellhead treating pressure. This pressure will be dictated by the client or
can be taken as some percentage of the maximum pressure rating of the surface
equipment, and will correspond to a given injection rate (that is, a given friction
pressure) through Eq. 1:

Pw = σ min + Pe − Ph + Pf (1)

Where:
pw = maximum allowable wellhead pressure (psi)
σmin = minimum in-situ stress (psi)
pe = fracture net pressure (psi)
ph = hydrostatic pressure (psi)
pf = total friction pressure (psi).
The fracture net pressure in acid fracturing is never larger than 250 psi (KGD or
RADIAL fracture geometries). The maximum injection rate can be calculated for any
acidizing fluid. Acid emulsions, ungelled acids and foamed ungelled acids exhibit
high friction pressure in the tubulars which may limit the maximum injection rate,
thereby limiting acid penetration into the fracture. Foams have another limiting
factor, low hydrostatic pressure, which can also limit injection rates.

2.5 Optimizing Conductivity and Etched Fracture Length


A unique feature of acid fracturing, as opposed to conventional fracturing with
proppant, is that there is no theoretical limitation to the conductivity of an etched
fracture. Increasing the conductivity of an acid-etched fracture is simply a matter of
pumping more acid. The more acid pumped, the wider the etched width in a
carbonate rock. Regardless of closure pressure and embedment strength of the
reservoir, increasing the etched width always means increasing fracture conductivity

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 21 of 37

after closure. This is opposite to the permeability of a propped fracture which is


related to proppant particle size and limited by proppant concentration.
A dimensionless fracture conductivity (CfD) value 50 or greater means that the
pressure drop in the etched fracture is negligible compared to the pressure drop in
the reservoir, which is the ideal case. Since

3
kfwf wf
FcD = = (2)
kx a 12kx a
Where:
xa = acid-etched fracture half-length (ft),
kf = acid-etched fracture permeability (md),
wf = after closure fracture width (in.) and
k = reservoir permeability (md),
the optimum fracture width after closure is well defined as soon as the maximum
acid penetration (xa) is known:
wf3 ≥ 600 k xa
The stimulation ratio is the ratio of the stimulated production rate to the initial
production rate at a given pressure differential between the wellbore and the
reservoir boundary. For a given fracture length, the maximum stimulation ratio that
can be achieved corresponds to the case of an infinite-conductivity fracture. A
reservoir with an infinite-conductivity fracture (CfD ≥ 50) is known to behave as a
homogeneous reservoir having an equivalent wellbore radius equal to one-half the
acid-etched fracture half-length, with a stimulation skin equal to:

Ss = ln( 2rw / x a ) (3)


Where:
Ss = stimulation skin (dimensionless)
rw = wellbore radius (ft).
This relationship is valid as long as the fracture half-length is small compared to the
drainage radius so that overall flow is approximately radial. This is normally the case
since acid etch-length is rarely greater than 200 ft. This negative skin can be used in
a Darcy flow equation to calculate the approximate reservoir productivity.

2.6 Effective Acid Concentration


The effective acid concentration is the effective strength of the acid taking into
account the nonacid components of the fluid. Gelled (or ungelled) 28% HCl has an
effective acid concentration of 28% while 15% HCl has an effective acid
concentration of 15%. However, an acid emulsion composed of 7 parts 28% HCl

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 22 of 37

and 3 parts diesel oil has an effective acid concentration of 19.6%. Similarly, a 70%-
quality foam having 15% HCl in its aqueous phase has an effective acid
concentration of only 4.5%. Larger volumes of these fluids must be pumped in order
to obtain a given fracture conductivity. If the acid is retarded, the retardation may
compensate somewhat for the low effective acid concentration. Foam also has the
additional drawback of providing low hydrostatic pressure. Except in areas with
cleanup problems, acid systems other than foams may have better application.

2.7 Selecting Fluids for Deeper Acid Penetration


To achieve fracture penetration, controlling the acid leakoff rate is imperative.
Reasons for acid leakoff in acid fracturing are
• natural fissures and fractures intersecting the hydraulic fracture
• wormhole growth into the fracture walls
• conventional fracturing fluid leakoff into porosity.
Decreasing Leakoff Through Natural Fissures
Adequate control of acid leakoff into fissures is not simple. The best approach is to
first bridge the inlet of the fissures with a fluid-loss additive such as a deformable
100-mesh resin, 100-mesh sand, or fine salt. The fluid-loss additive can serve as an
induced (artificial) rock matrix and base for a wall-building fluid-loss additive. The
fluid-loss additive used for bridging should contain particles that are smaller in
diameter than the average fissure width but still large enough to support bridging.
Estimates of the fissure-width and spacing are possible by the examination of
formation cores or by logging techniques such as the Formation MicroScanner*,
BHTV and SDT logs (Schlumberger Wireline and Testing).
LCA fluids and viscous pads can also be used to decrease leakoff through natural
fissures.
The effectiveness of a second fluid-loss additive, the wall-building fluid-loss additive,
in sealing the artificial matrix in the fissure inlets will be greatly enhanced if the
individual particles will deform and adhere to each other. Thus the selection of the
appropriate fluid-loss additive or fluid-loss additive blend is critical. The Fracturing
Materials Manual  Additives, Matrix Materials Manual and the Dowell publication
Fluid Loss Additives and Diverting Agents (TSL-2064) should be referenced before
selecting the types of fluid-loss additives to use.
Fluid-loss additives should be added to gelled fluid stages that precede and are
interspersed with the acid stages. The presence of viscous fluid within the bridged
fissure will also contribute to fluid-loss control.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 23 of 37

Decreasing Acid Leakoff Due to Wormholes


There are several different methods of reducing acid leakoff due to wormholing.
These are (from most to least efficient)
• use LCA fluids
• fill existing wormholes with viscous, non-reactive fluids
• viscosify the acid.
LCA fluids combine the advantages of having high viscosity plus a unique property of
temporarily crosslinking at the interface between the acid and the formation. The
crosslinked interface provides some of the desirable properties found in a
conventional wall-building fluid-loss additive. Additional information on this process
is provided in the Fracturing Materials Manual — Fluids.
Viscosifying (gelling) the acid decreases the leakoff rate according to Darcy's law,
that is, for a constant pressure differential, increasing the viscosity decreases the
flow rate. This, in turn, decreases wormhole length.
Some acid viscosifiers maintain a high viscosity while spending, but break when the
acid is spent. Additional information on gelled acids is provided in the Fracturing
Materials Manual — Fluids.
The fracture acidizing treatment technique, DUOFRAC II, is designed to stop
wormhole growth once growth has begun. The technique consists of initiating a
fracture with a nonreactive pad of gelled fluid followed by alternating stages of acid
and nonreactive pad. The first pad volume initiates a fracture for the first acid stage
to follow. The first acid stage etches a portion of the fracture face and also creates
leakoff wormholes that must be controlled. The second pad stage fills the initiated
wormholes and prevents the second acid stage from entering the established
wormholes. This acid stage will etch the next increment of fracture as well as create
new wormholes in the next fracture increment. The process is repeated until the
designed treatment volumes are depleted. The higher the viscosity of the
nonreactive pad fluids, the more difficult it is for acid to displace the gelled fluid and
resume wormhole growth.
Decreasing Leakoff Through Fracture Walls
In proppant fracturing, the most common way to control fluid loss is to build an
impermeable filter cake on the fracture walls. In acid fracturing, viscosity is much
more effective for fluid-loss control.
As an alternative, if the first stage of an acid fracturing treatment is a linear, inert fluid
(with no wall-building characteristics) some leakoff control will be provided. The fluid
will leak off into the native porosity of the fracture faces and will form a layer of
viscous fluid in the first several inches of reservoir rock adjacent to the fracture. The
viscous fluid bank will reduce acid leakoff better than a conventional polymer cake
that does not offer any resistance to acid.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 24 of 37

The viscous fluid bank functions to control leakoff as a viscosity-controlled fluid. This
approach to reducing leakoff is of benefit only in reservoirs where the resulting
viscosity-controlled leakoff coefficient is smaller than the compressibility-controlled
leakoff coefficient (for example, gas wells).
Finding the correct information for this design may be difficult due to a lack of
information about true leakoff viscosities of linear gels. Conventional American
Petroleum Institute (API) fluid-loss tests are almost always interpreted in terms of
their wall-building coefficient with the assumption that the leakoff viscosity is the
viscosity of the base fluid.
This design should not be used in wells with low bottomhole pressure because
cleanup problems may occur. In this case, another technique can be used, resulting
in a decrease of the apparent reservoir permeability along the fracture faces. This is
achieved by emulsifying or foaming the treatment fluids. The leakoff is reduced as a
result of the two-phase flow effects of the treating fluid leaking off into the reservoir
rock adjacent to the fracture. Cleanup is expected to be easier, especially when the
overflush stage incorporates large amounts of mutual solvents, surfactants and/or
demulsifiers.

2.8 Determination of Leakoff Coefficients


Laboratory studies of acid fracturing systems cannot derive leakoff coefficients
representative of those that occur during acid fracturing when wormholing takes
place. In the laboratory, a wormhole is induced and penetrates a small diameter
core. The effect of leakoff through this induced wormhole is measured as a function
of the cross-sectional area of the small diameter core. However, in the reservoir,
wormhole density (the number of wormholes per unit of leakoff area) is thought to be
considerably less. As a consequence, laboratory leakoff values calculated from
artificially induced wormholes are (probably) greater than downhole conditions.
These laboratory tests are, however, useful to compare the relative leakoff data and
etch characteristics of the various acid systems. The following methodology has
been derived using relative leakoff data from the laboratory.

2.8.1 Methodology
1. Estimate the apparent reservoir permeability (ka) during the acid stages.
Based on the considerations discussed previously (Acid Leakoff), the matrix
permeability (kr) of the reservoir is estimated to be approximately five percent of
the total effective permeability of a fissured reservoir.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 25 of 37

The apparent permeability can be estimated along the following guidelines.


• If the reservoir is not fissured, ka = kr
• If the reservoir is fissured, and if
√ no fluid-loss additive is used, ka = kr
√ a fluid-loss additive is used in linear pads before each acid stage:
k
ka = r
3

√ a fluid-loss additive is used in crosslinked pads before all acid stages, or


k
LCA fluids are the acid system, ka = r
10

2. In Fig. 7, Table 3, Table 4, and Table 5, the permeability to consider is the


apparent permeability and the temperature is the average temperature in the
fracture (Ta).

Fig. 7. Viscosity of fresh water versus temperature.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 26 of 37

Table 3. Total Minimum Leakoff Coefficients versus Temperature and


Permeability
Comparative values for Ctmin can be obtained using the table below. The
numbers shown were compiled from Cw and spurt values for crosslinked
fluids provided in the Fracturing Materials Manual — Fluids . Since these are
estimated values, extrapolations for other permeabilities and temperatures
can be made.
Total Leak-Off Coefficient (Ctmin = ft/min½)

Apparent Temperature
Permeability °F 100 150 200 250
(md) °C 38 66 93 121
<0.1 0.0005 0.0005 0.0010 0.0015
1 0.0010 0.0015 0.0020 0.0030
>10 0.0025 0.0030 0.0040 0.0055
3. Determine the viscosity, compressibility coefficient (Cvc) for water, using of the
apparent permeability and the other reservoir data.
Water viscosity taken at Ta is used to estimate the maximum leakoff coefficient,
Ctmax, of an inert fluid in the studied reservoir (that is, the poorest control of fluid
loss).
Eq. 3, Eq. 4, and Eq. 5 in Appendix E — Fluid Loss can be used instead of the
FracCADE software to obtain the Cvc for water, that is, Ctmax. Similar results should
be obtained providing that an average fracture net pressure of 250 psi is
considered, and that identical compressibility and saturation values for rock, oil,
gas and water are used.
4. Determine Ctmin, the minimum leak-off coefficient for an inert fluid (best leakoff
control) from Table 3.
5. The value for Ctmin should not be greater than that of Ctmax. If it is, review the
reservoir data and correct. When the values are confirmed, then the total
coefficient values CtF for all fluids will be equal to Ctmax.
6. Determine the scale of efficiency (SFb) for controlling leakoff using Table 4.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 27 of 37

Table 4. Scale Factor Base Values (SFb)


This table below, based on comparative laboratory test results, presents the degree of
leakoff control expressed as values between zero (best control) and eleven (poorest
control) for reactive and nonreactive fluids.
Limestone °F/°C ≤100/38 ≤150/66 ≤200/93 ≤250/121 >250/121
Fluid Dolomite °F/°C ≤200/93 ≤250/121 ≤300/149 ≤350/177 >350/177
Ungelled Acid + F98 NA NA NA 11 10
Ungelled Acid 8 11 10 9 8
DGA*200, 300, and 400 5 8 7 6 5
SXE fluids NA NA NA 8 7
LCA fluids 3 4 4 3 3
Water ----------------------------------------------8-------------------------------
--
Linear Pad ----------------------------------------------3-------------------------------
--
Crosslinked Pad ----------------------------------------------0-------------------------------
--
NOTE
a) For ungelled acid, SFb should be decreased by a value of one for each 7% decrease
in acid concentration below 28%.
b) For the other acid systems, viscosity tends to dampen decreases in acid
concentration.
c) SFb values for organic acid are given the same values as used with ungelled 15%
HCI + F98.
d) For all fluids:  Decrease SFb by one if energized,
 Decrease SFb by two if foamed.

7. Using the previously determined SFb values, corrected scale factors, (SFc), should
now be derived according to the guidelines in Table 5. Corrections are
necessary to account for the effects the different fluid systems have on each
other.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 28 of 37

Table 5. Scale Factor Corrections


These corrections are necessary to account for the effects the different fluid
systems have on each other.
REACTIVE FLUIDS THE MINIMUM SFc VALUE IS ONE
a) Decrease SFb by one when the apparent permeability is less than 1 md;
and again by one for each order of magnitude ten decrease of the
apparent permeability. This is reflective of limited wormhole propagation
due to small pore radii.
b) After the first pad of linear gel, decrease SFb by two to account for viscosity
controlled leak-off.
c) After the first pad of crosslinked gel, decrease SFb by one. Perforation of
the polymer cake will be rapid.
d) Decrease SFb by two after other pads, linear or crosslinked. This value is a
tradeoff between wormhole filling at the fracture inlet where crosslinked
fluid is better and on new fracture face at the fracture tip where linear fluid
is better.
NONREACTIVE FLUIDS  THE MINIMUM SFc VALUE IS ZERO
a) Increase SFb by one when the nonreactive fluid follows an acid stage
having a SFb value between three and six,
b) Increase SFb by two when the non-reactive fluid follows an acid stage
having a SFb value between six and nine,
c) Increase SFb by three when the non-reactive fluid follows an acid stage
having a SFb value of greater than nine. This accounts for the spurt that
takes place when wormholes are filled by the gelled non-reactive fluid.

8. The final total leakoff coefficient (CtF) of the fluid being considered is determined
using Eq. 4.

CtF = Ct min + 0.125 × (Ct max − Ct min ) × SFc (4)


Where:
CtF = total leakoff coefficient (ft/min½ ),
Ctmin = minimum leakoff coefficient (ft/min½), and
Ctmax = maximum leakoff coefficient (ft/min½).
When using Eq. 4, note that for:
SFc = 0, CtF = Ctmin. (case for a crosslinked pad),
SFc = 8, CtF = Ctmax. (case for a plain-water pad),

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 29 of 37

Note also that some stages can have larger leakoff coefficients (even poorer fluid
leakoff control) than plain water if their SFc is greater than 8.
CtF values determined in the above manner must be used in conjunction with zero
spurt and with leakoff coefficient flag on TOTAL in the Fluid Editor of FracCADE.
These corrections are necessary to account for the effects the different fluid systems
have on each other.

2.8.2 Example Calculation


Reservoir Parameters:
• fissured dolomite oil reservoir
• permeability 10 md, porosity 12%
• oil viscosity = 0.6 cp
• total compressibility = 5 X 10-5 psi-1
• bottomhole static pressure = 1770 psi
• minimum in-situ stress = 5120 psi
• fracture net pressure = 0 psi
• average temperature in the fracture = 200°F

Pumping schedule (using DUOFRAC II system):

1) Crosslinked pad with fluid-loss additive


2) 20% HCl
3) Crosslinked pad
4) 20% HCl
5) Water overflush
1. Estimate the apparent permeability (ka) of the reservoir during the acid stages:
Fissured reservoir, but a crosslinked pad with fluid-loss additive is used:
ka = kr /10 = 1 md
2. Verify ka and Ta.
ka = 1 md, Ta = 200°F

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 30 of 37

3. Determine Cvc for water at Ta = 200°F.


Water viscosity at 200°F = 0.3 cp
1/ 2
 1 × ( 5120 − 1770 + 0 ) × 0.12 
Cν = 1.48 × 10 − 3  
 0.3

= 0.0542 ft / min 1/ 2
1/ 2
−3  1 × 5 × 10 − 5 × 0.12 
Cc = 119
. × 10 ( 5120 − 1770 + 0)  
 0.6 

= 0.0126 ft / min 1/ 2

2 × 0.0542 × 0.0126
Cνc =
[ ]
1/ 2
0.0542 + (0.0542) 2 + 4( 0.0126 ) 2

Cνc = 0.0120 ft / min 1/ 2


[Note that Cvc for water in the case of no control of leakoff into natural fissures
(ka = kr= 10 md) would be equal to 0.0379 ft/min½, that is, the Cvc above times:
10 ].
4. Determine the minimum leak-off coefficient for a nonreactive pad.
At Ta = 200°F and ka = 1 md and using ,
Ctmin = 0.0020 ft/min½
5. Verify that Ctmin is less that Ctmax.
Ctmin = 0.0020 ft/min½, Ctmax = 0.0120 ft/min½.
6. Determine the SFb values using Table 4. See pumping schedule.
a. Crosslinked pads. SFb for crosslinked pad = 0
b. 20% HCl, 200°F. SFb for 20% HCl = 7 (Adjusted 8 - 1 = 7 for acid
concentration)
c. Water Overflush. SFb for Water Overflush = 8
7. Determine SFc values using Table 5. Follow the pumping schedule.
a. Crosslinked pad. SFc = 0 (No correction.)
b. 20% HCl. SFc = 6 (Corrected 7 - 1, Acid follows crosslinked pad.)
c. Crosslinked pad. SFc = 2 (Corrected 0 + 2, second pad stage follows the first
acid stage.)
d. 20% HCl. SFc = 5 (Corrected 7 - 2, Second stage of acid follows second stage
of pad.)

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 31 of 37

e. Water Overflush. SFc = 9 (Corrected 8 + 1, follows second acid stage having


an SFc of 5.)
8. Determine the CtF for each fluid using .
a. Crosslinked pad (first one) SFc = 0, CtF = 0.0020 ft/min½
b. 20% HCl (first stage) SFc = 6, CtF = 0.0095 ft/min½
c. Crosslinked pad (second one) SFc = 2, CtF = 0.0045 ft/min½
d. 20% HCl (second stage) SFc = 5, CtF = 0.01075 ft/min½
e. Water overflush SFc = 9, CtF = 0.01325 ft/min½

2.8.3 Notes
When estimating CtF for any stage of any fluid for a particular treatment, everybody
has to exercise their own judgment. For instance, not using any fluid-loss agent in a
second pad stage obviously should impair the leakoff control abilities of the
considered pad and following stages. A good indication on how much this effect
would be is given by recalculating the Ctmin and Ctmax with ka = kr. However, this loss of
efficiency only concerns the tip of the fracture; therefore, CtF have to be worked out
as averages between the fully controlled case (with ka = kr/10) and the fully
uncontrolled one (with ka = kr).
The “wormhole perimeter” coefficient in the acid simulator screen of FracCADE
(SETUP option REGULAR) is a research tool which must not be used when
designing actual jobs and using the method of determination of leakoff coefficients
described above. The default value of 0 in the FracCADE software actually means
that this coefficient is not considered in the estimation of the leakoff volumes by the
simulator. Any other values would have to be used in conjunction with actual leakoff
data generated in the laboratory on one-in. diameter limestone cores.

2.9 Cooldown
Temperature controls both the diffusion and the surface-reaction rates. At high
temperatures limestones and dolomites react very rapidly with hydrochloric acid; so
rapidly, that acid-etching of the fracture may be limited to only a few feet. If acid
fracturing is the only viable means of fracture stimulation, cool-down must be
considered.
The best fluids for cooldown are high-leakoff fluids. An example would be a water-
base fluid containing a friction reducer. Obviously, the larger the volume of cold fluid
invading the formation and the higher the injection rate, the lower the cooldown
temperature in the fracture and the longer it takes the reservoir temperature to
rebound and heat the fracture to the original bottomhole static temperature when
leakoff stops. The most important aspect of fracture cooldown is the invasion of the
primary porosity of the reservoir rock adjacent to the fracture faces.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 32 of 37

Injecting large amounts of a cold fluid into a pattern of natural fissures will not
succeed at notably decreasing the temperature of the rock between the fissures.
Therefore, the first requirement in a fissured reservoir is to consider limiting leak-off
into the fissures. Once the leakoff to fissures has been limited, cooldown can be
initiated. Temperature models show that the temperature in a fracture induced by a
low-efficiency fluid such as water has a linear profile. At any time during the
propagation of the fracture, the temperature at any distance from the wellbore can
be calculated using Eq. 5.

[T ( X ) − Tw ] = [Tbhs − Tw ] (5)
x xf
Where
T(X) = fluid temperature in the fracture at distance (X) from the wellbore
(°F [°C])
Tw = wellbore temperature at the perforations (°F [°C])
X = distance from the wellbore (ft)
Tbhs = bottomhole static temperature (°F [°C])
xf = hydraulic fracture half-length (ft).
In most cases, Tw is reached rapidly and levels-off at a value slightly above the
surface temperature of the cooldown fluid (normally from 10° to 30°F [5° to 15°C]
above the surface temperature of the cooldown fluid). This depends on the injection
rate and the volume injected. Circulating the wellbore with cold water before an acid
fracturing treatment (before initiating formation cooldown) will allow for more
confidence when estimating Tw.

Calculating Cooldown
A simple method of designing a cooldown treatment is:
1. Determine the maximum acid penetration (xa) at a cooldown temperature (Ta)
using the FracCADE software.
2. Determine the required hydraulic fracture half-length at the end of cooldown in
order to obtain (Ta) at the distance (xa) from the wellbore (Eq. 6).
Tbhs − Tw
x f = xa
Ta − Tw (6)
Where:
xa = acid-etched fracture half-length (ft)
Ta = fluid temperature at distance (xa) from the wellbore (°F [°C]).
3. Run the FracCADE software again to determine the required cooldown pad
volume to achieve (xf) before acid injection.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 33 of 37

2.10 Retarded Acid


A retarded acid is an acid with a reduced reaction rate. The degree of retardation is
defined by a retardation factor. The retardation factor for plain HCl at any
concentration is one. Therefore, a retarded acid has a retardation factor greater
than one. Most retarded acids are modifications of HCl. When compared to
unretarded HCl, retarded acid penetrates more deeply into the hydraulic fracture
thereby increasing acid-etched length. If all other factors remain unchanged,
fracture width is decreased because the acid has been altered so that it can
penetrate deeper into the hydraulic fracture before becoming spent. The retardation
factor is derived by comparing the overall reaction rate of a retarded acid to the
overall reaction rate of nonretarded hydrochloric acid. Specifically, it is the ratio of
the slope of the spending profile for the retarded acid to the slope of the spending
profile for nonretarded HCl under an identical set of conditions. provides retardation
factors for various acid systems and guidelines for modifying the retardation factors.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 34 of 37

Table 6. Retardation Factor Selection Guidelines


The values shown are guidelines for selecting retardation factors (RF). Note
that small deviations from the retardation factor value for plain HCl will have
large effect on the acid penetration calculated by the FracCADE software.
1. Base Values
RF
a. HCl, DGA and LCA systems 1.0
b. Acid-external emulsions (DAD) 2.0
c. Surfactant retarded with F98 2.0
d. Organic acids (formic or acetic) 4.0
e. Acid-internal emulsions (SXE) 10.0
2. Modifications to the Base Values
Add to RF
a. If linear pad before acid stage 0.5
b. If crosslinked pad before acid stage 1.0
c. If acid stage energized with CO2 0.5
d. If acid stage foamed with CO2 1.0
e. If acid stage foamed with N2 only 0.5

Determining the Retardation Factor by Dynamic Testing


The retardation factor must be measured under dynamic conditions. Under static
conditions, for example, the spending times of gelled acids are extended (Fig. 8).
The gelled acid only appears retarded. Under dynamic conditions, the spending time
of Gelled Acid A is the same as 15% HCl (Fig. 9). Under static conditions, Gelled
Acid A appears retarded because its viscosity prevents immediate dispersion of acid
reaction products (CO2, CaCl2 and water). These reaction products form a
temporary barrier between the carbonate rock and live acid. Under dynamic
conditions, these reaction products are swept away and the retardation factor of one
is maintained.
This does not imply that every acid containing a surfactant is retarded nor is every
gelled acid not retarded. Each acid solution must be tested under the same dynamic
conditions before meaningful comparisons can be made.
It is valid, in a narrow sense, to consider gelled acids that utilize no retarding
mechanism to be retarded. The viscosity of a gelled acid will provide a wider
fracture than ungelled acid (and sometimes will switch flow conditions from turbulent
to laminar). The viscosified acid will penetrate deeper into the reservoir before

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 35 of 37

convection causes contact with the fracture faces. The FracCADE software
accounts for these factors.
Fig. 8 and Fig. 9 compare the spending time of acid and limestone in static tests
versus the retardation factor of the same acid systems under dynamic conditions.
These include acid concentration, temperature, pressure, rheology, type of
carbonate rock used, fluid saturation of the rock, injection rate unit of fracture height
and initial fracture width.
Static tests are of little value when determining the behavior of an acid system in a
reservoir.

Fig. 8. Spending time versus HCl concentration at static conditions.


Note that Gelled Acid A and Surfactant Retarded Acid B appear retarded.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Acid Fracturing Dowell
Page 36 of 37

Fig. 9. Temperature versus retardation factor at dynamic test conditions.

2.11 Viscous Fingering


Viscous fingering is a physical phenomenon that increases acid penetration.
Fingering occurs when a viscous fluid is displaced by a less viscous fluid; the
interface of the two fluids being very unstable. Pumping acid behind a more viscous
pad will cause multiple fingers of acid to quickly extend into the pad.
Acid fingering has three positive effects.
1. The overall velocity of the acid in the fracture is increased slightly by the fingering
behavior.
2. The acid-etched length is increased.
3. The acid leakoff area is decreased.
A DUOFRAC II treatment will also experience viscous fingering, but this fingering is
probably secondary to the original purpose of the DUOFRAC II procedure. The
primary intent is to increase hydraulic fracture width and to provide acid leakoff
control.

DOWELL CONFIDENTIAL
Section 900
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Acid Fracturing
Page 37 of 37

The FracCADE software does not model viscous fingering behavior, rather
accounting for fluid displacement in a fracture using plug-flow calculations. Modeling
of viscous fingering would require determination of the number, height and width of
acid fingers in the fractures and accurate vertical permeability measurements. The
value for each of these properties is highly subjective.

2.12 Summary of Treatment Design Fundamentals for Acid Fracturing


Fluid selection is relatively easy if the objective of the acid fracturing treatment is to
maximize the acid penetration.
• Always use the maximum true acid concentration.
• Always consider acid leakoff. Consider the DUOFRAC II process or low-leakoff
systems or both.
• Select pad fluids and acid systems that will create deep flow channels within the
fracture.
• Temperature (ranges below are for chalks and limestones  add 100°F for
dolomites):
− Less than 100°F (38°C): Use straight acid at lowest injection rate (though
above fracturing rate) to compensate for very slow reaction rate.
− 100° to 150°F (38° to 66°C): Candidate acid systems are gelled acids and
low-leakoff acids.
− 150° to 200°F (66° to 93°C): Candidate systems are high-temperature
gelled acids or low-leakoff acids.
− 200° to 250°F (93° to 121°C): Candidate systems are acid-in-oil emulsions
and high temperature gelled acids.
− Greater than 250°F (121°C): Consider cooldown procedures or retarded
acids.
• Retarded acid can increase etched fracture penetration.
Energizing with CO2 or N2 may be necessary in wells where cleanup problems are
anticipated. Foaming is not recommended where deep penetration is required
because the effective acid concentration is reduced significantly and injection rate is
limited.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 1 of 43

HORIZONTAL WELLS

1 Introductory Summary............................................................................................................. 4
1.1 Candidate Reservoirs for Horizontal Wells .......................................................................... 4
1.1.1 Naturally-Fractured Reservoirs................................................................................... 4
1.1.2 Matrix-Permeability Reservoirs................................................................................... 4
1.1.2.1 Vertical Permeability ........................................................................................ 5
1.1.2.2 Skin Damage ................................................................................................... 6
1.2 Fracture Performance and Wellbore Orientation ................................................................. 7

2 Fracture Orientation ................................................................................................................ 8

3 Fractured Horizontal Well Performance............................................................................... 10


3.1 Longitudinal Fracture ......................................................................................................... 10
3.2 Orthogonal Fractures ......................................................................................................... 11
3.3 Choke Skin Effect............................................................................................................... 13
3.3.1 Fracture Reorientation — Choke Effect.................................................................... 14
3.4 Coning Effects.................................................................................................................... 15
3.4.1 Comparison of Fractured and Nonfractured Reservoir............................................. 15
3.4.2 Effect of the Distance From the Fracture to the Water Zone.................................... 15

4 Rock Mechanical Properties ................................................................................................. 17


4.1 Openhole Wellbore Stability............................................................................................... 17
4.2 Shear Failure ..................................................................................................................... 18
4.3 Tensile Failure ................................................................................................................... 18
4.4 Matrix Collapse .................................................................................................................. 18
4.5 Cased-Hole Wellbore Stability ........................................................................................... 19
4.6 Stress and Deformation Analysis ....................................................................................... 20

5 Fracture Initiation And Propagation..................................................................................... 20


5.1 Initiation Pressure .............................................................................................................. 20
5.2 Fracture Initiation ............................................................................................................... 21

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 2 of 43

5.3 Fracture Propagation..........................................................................................................23


5.4 Longitudinal Fractures ........................................................................................................23
5.5 Angled Fractures ................................................................................................................24
5.6 Transverse Fractures .........................................................................................................25
5.7 Controlling Fracture Reorientation......................................................................................26

6 Perforating ..............................................................................................................................27

7 Treatment Design ...................................................................................................................29


7.1 Net Present Value Analysis ................................................................................................29
7.1.1 Calculating the NPV of Orthogonal Fractures ...........................................................29
7.1.2 Horizontal Well Production Prediction .......................................................................31
7.2 Fracture Height...................................................................................................................34
7.3 Fracture Orientation ...........................................................................................................34
7.4 Fracture Length and Conductivity.......................................................................................34
7.5 Pump Rate..........................................................................................................................34
7.6 Fracturing Fluid Selection...................................................................................................34
7.7 Proppant Selection .............................................................................................................35
7.7.1 Mesh Range..............................................................................................................35
7.7.2 Proppant Type...........................................................................................................35
7.7.3 Proppant Concentration ............................................................................................35

8 Execution ................................................................................................................................35
8.1 Perforating ..........................................................................................................................35
8.1.1 When to Perforate .....................................................................................................36
8.2 Wellbore Isolation Between Fractures................................................................................36
8.2.1 Isolation Using Mechanical Tools..............................................................................37
8.2.2 Isolation Using Proppant Plugs .................................................................................37
8.2.2.1 Intentional Screenout .....................................................................................37
8.2.2.2 Multidensity/Multimesh Proppant ...................................................................38
8.2.3 Isolation Using Viscous Plugs ...................................................................................38
8.2.4 Wellbore Isolation in an Openhole Completion .........................................................39
8.3 Flowback ............................................................................................................................39

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 3 of 43

9 Overview of the Horizontal Well Treatment-Design Procedure ......................................... 40

10 Case History ......................................................................................................................... 41

FIGURES

Fig. 1. Productivity index ratios for horizontal versus vertical wells. ............................................ 6
Fig. 2. A longitudinal fracture. ...................................................................................................... 8
Fig. 3. An orthogonal fracture. ..................................................................................................... 9
Fig. 4. A fracture propagating at an angle to the wellbore. .......................................................... 9
Fig. 5. Productivity index ratios of vertical well/vertical fracture and horizontal well with a
longitudinal fracture......................................................................................................... 10
Fig. 6. NPV analysis for a vertical well. ...................................................................................... 12
Fig. 7. NPV analysis for one orthogonal fracture. ...................................................................... 12
Fig. 8. Water breakthrough time versus the position of the fracture with respect to
the oil/water contact. ....................................................................................................... 16
Fig. 9. Production rate versus water-cut. ................................................................................... 16
Fig. 10. Fracture conductivity versus water-cut. ........................................................................ 17
Fig. 11. Mohr failure envelope for matrix collapse. .................................................................... 18
Fig. 12. Stress distribution around a perforation. ....................................................................... 20
Fig. 13. Initiation pressure as a function of α and the borehole inclination. ............................... 21
Fig. 14. Horizontal well configuration in the in-situ stress field................................................... 21
Fig. 15. Initiation points and fracture orientation on the borehole. ............................................. 22
Fig. 16. Fracture initiation pressure. .......................................................................................... 22
Fig. 17. The effect of distance between collinear fractures on maximum fracture width............ 24
Fig. 18. Fracture rotation angle versus spacing......................................................................... 25
Fig. 19. Width and excess pressure as a function of spacing for parallel, transverse
and radial fractures........................................................................................................ 26
Fig. 20. Radius of fracture reorientation as a function of the ratio between the maximum
and minimum horizontal stresses. ................................................................................. 27
Fig. 21. Critical distance between perforation versus well orientation. ...................................... 28
Fig. 22. Single phase flow .......................................................................................................... 33
Fig. 23. NPV analysis of the number of orthogonal fractures..................................................... 42
Fig. 24. Actual versus predicted fluid production. ...................................................................... 43

TABLES

Table 1. Performance Comparison Of Vertical And Horizontal Wells With Fractures................ 13


Table 2. Pseudoskin Factor Correlation Contstants................................................................... 32
Table 3. Overview of Fracture Treatment Design Considerations for Horizontal Wells ............. 40
Table 4. Stress Measurement Techniques ................................................................................ 41

1 Introductory Summary
Successful hydraulic fracturing of wells with horizontal or deviated wellbores
depends upon a number of factors. The most important factors are

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 4 of 43

• The fractured well should be a viable economic alternative to a fractured vertical


well or wells.
• Stability risks, such as wellbore or casing collapse during drilling, stimulation or
production, are understood and are considered acceptable.
• Cementing or completion or both is satisfactorily accomplished.
• A decision to create a single hydraulic fracture, or to create multiple fractures
simultaneously or in stages has been made.
Horizontal wells that are candidates for hydraulic fracturing are usually those with
low natural permeability, unfavorable permeability anisotropy and no natural
fractures.

1.1 Candidate Reservoirs for Horizontal Wells


The most prevalent economic incentive is the increased productivity of a horizontal
well as opposed to a fractured vertical well. In many cases, horizontal wells are
drilled specifically to minimize water and gas coning rather than to increase oil or gas
production.

1.1.1 Naturally-Fractured Reservoirs


The benefits of a horizontal well (versus a vertical well) in a naturally-fractured
reservoir are
• an increased number of natural fractures are intersected (production is
increased)
• drainage area is increased
• the need for hydraulic fracturing is eliminated.
The most successful horizontal wells (and the majority of horizontal wells) are
completed in naturally-fractured reservoirs.

1.1.2 Matrix-Permeability Reservoirs


The benefits of a horizontal well (versus a vertical well) in a matrix permeability
reservoir are
• reservoir contact with the wellbore is increased (increased productivity index)
• linear flow in the reservoir (reduced water and gas coning)
• drainage area is increased.
A productivity increase in a matrix permeability reservoirs is a result of an increase in
reservoir area (length) in contact with the wellbore. The reservoir contact in a
vertical well is limited to the height of the reservoir. The most benefit from this
approach is in thin reservoirs or in reservoirs with a thin oil column. In the case of a
thin oil column, a horizontal well has the additional advantage of increased

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 5 of 43

production with less drawdown, that in many cases will reduce or eliminate water or
gas coning.
There is less tendency for water or gas coning to occur in a horizontal well because
a horizontal well requires a much smaller pressure drawdown to produce at the
same rate as a vertical well. The pressure gradient from the horizontal wellbore to
the drainage radius is almost linear as opposed to a vertical well where there is a
log-linear pressure gradient from the wellbore to the drainage radius. This results in
a steady gas dip or water crest over the producing length of the horizontal wellbore
rather than a cone flowing to a single point source (which accelerates water or gas
coning) as in a vertical well.
A productivity increase is not simply equal to the ratio of the length of the horizontal
well to the thickness of the reservoir. Vertical permeability and skin damage are
limiting factors to well performance.

1.1.2.1 Vertical Permeability


The negative impact of low vertical permeability compared to horizontal permeability
is illustrated in where β = (kh/kv)½.
Fig. 1 shows a comparison of productivity index (PI) ratios (q/∆p) for three situations;
common anisotropy (β = 3), complete isotropy (β = 1), and highly favorable vertical
anisotropy (β = 0.25). The comparison is made for three net heights; 20 ft, 100 ft,
and 200 ft. For this illustration, the drainage radius is 0.326 ft (7.875-in. borehole).
Fig. 1 indicates that the value of β is crucial. If β is typical (≈ 3 for common
anisotropy), the PI ratio will be small. Conversely, if β is small, as would be the case
in formations with massive natural vertical fissures, then the PI ratio can be
extremely large and a horizontal well is an obvious choice.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 6 of 43

Fig. 1. Productivity index ratios for horizontal versus vertical wells.


Fig. 1 also indicates that the reservoir thickness is important. Horizontal wells are
less attractive for thicker formations and the PI ratios decrease as the formation
becomes thicker. The majority of horizontal wells are drilled in formations less than
150-ft thick.
Inverted high-angle wells, stepped wells and hydraulic fracturing (not a simple
solution) can offset low vertical permeability.
Water and gas coning can be less readily controlled in wells with high vertical
permeability. The extreme case is when small faults or natural fractures are present
in the formation.

1.1.2.2 Skin Damage


The presence of skin damage in a horizontal well can be determined by type-curve
matching or with the STAR* software. Most horizontal wells drilled in matrix
permeability are damaged. Cuttings that are ground into a fine paste by the
movement of the drill string in the horizontal wellbore will plug the pore throats in the
matrix close to the wellbore. This is more of a problem in horizontal wells than
vertical wells because the drill string (in a horizontal wellbore) lies on the low side of
the wellbore and PDC bits (often used to drill a horizontal well) produce fine cuttings
(as opposed to roller bits that produce coarse cuttings).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 7 of 43

1.2 Fracture Performance and Wellbore Orientation


The behavior and performance of a fracture propagating from a horizontal wellbore
is a result of the wellbore orientation with respect to the principal stresses in the
formation.
σHmin)
Wellbore in the Direction of the Minimum Horizontal Stress (σ
• Production from the fracture is reduced by limited contact with the wellbore
because of the choke skin effect (Sch)c.
• Multiple orthogonal fractures are required to maximize the productivity of the well
in terms of net present value (NPV).
• Higher breakdown pressure than a vertical well.
• A short perforated interval (less than four times the wellbore diameter) is required
to prevent the creation of multiple fractures from the same interval.
• Fracture spacing must exceed one-half the fracture half-length to prevent
interference (reduced fracture width, increased net pressure and rotation).
Wellbore in the Direction of the Maximum Horizontal Stress (σ σHmax)
• Productivity from the fracture is not choked by limited contact with the wellbore.
• The equivalent of a single fracture can be created along the wellbore.
• Lower breakdown pressure than a vertical well.
• Little interference between fractures.
• Length of the perforated interval may not be so critical.
Wellbore at an Angle Between σHmin and σHmax
• Production is limited by the effects of choke skin effect and additional choke
effect (Sfs)c.
• Multiple fractures are reoriented at some point away from the fracture.
• Fractures have nonidentical wings.
• Higher friction pressure and possible proppant bridging because of fracture
reorientation.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 8 of 43

2 Fracture Orientation
The orientation of hydraulic fractures in a horizontal well depends on the magnitude
and orientation of in-situ stresses. Because of the stresses being concentrated near
the wellbore, the direction in which a fracture initiates at the wellbore will seldom be
the direction in which it will ultimately propagate, which is perpendicular to the
direction of the minimum principal stress, or along a dominant natural fracture.
Therefore, in many cases, hydraulic fractures may not be planar features.
A fracture which initiates and then propagates longitudinally along the axis of a
wellbore will do so from a number of perforations at the wellbore (Fig. 2). Therefore,
it is reasonable to presume that the pressure drop, while treating or producing the
well, will be similar to that of a vertical well.

Fig. 2. A longitudinal fracture.


A fracture which initiates or propagates or both, at an angle to the wellbore will be in
communication with fewer perforations. When the hydraulic fractures are orthogonal
to the direction of the wellbore (Fig. 3), the inefficient contact between the fracture
and the well can be quantified by introducing a choke skin effect (Eq. 1).

( Sch )c =
kh
kfw
[1n(h / 2rw ) − π / 2] (1)
Where:
(Sch)c = choke skin effect (dimensionless)
kh = horizontal permeability (md)
kfw = fracture conductivity (md.ft)
h = net reservoir thickness (ft)
rw = wellbore radius (in.).

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 9 of 43

Fig. 3. An orthogonal fracture.


Fractures which propagate in a direction between the two extremes will have a much
more complicated geometric shape (Fig. 4).
A fracture may initiate in a longitudinal direction and then reorientate so that it grows
perpendicular to the minimum principal stress, once is has grown away from the
influence of the wellbore and the completion. In this case, the maximum fracture
width may not occur at the wellbore and may not vary uniformly along the length of
the fracture. The reorientation of the fracture and its nonuniform geometry will act in
some degree to limit production from the fracture.

Fig. 4. A fracture propagating at an angle to the wellbore.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 10 of 43

3 Fractured Horizontal Well Performance


When planning to fracture an existing horizontal well or to drill a well which will
require fracturing, a careful study must be made to determine the orientation of the
fracture or fractures and how this may impact any increased productivity. In many
cases, fracturing a horizontal well will further delay the onset of water or gas coning.

3.1 Longitudinal Fracture


A comparison of a vertical fractured well with a longitudinally fractured horizontal well
is shown in Fig. 5.

Fig. 5. Productivity index ratios of vertical well/vertical fracture and


horizontal well with a longitudinal fracture.
Productivity index ratios of vertical well/vertical fracture and horizontal well with a
longitudinal fracture. This comparison is for a fully penetrating fracture (that is,
2xf = L) and a fractured vertical well with a fracture half-length equal to xf. The CfD
values for the fractured well are graphed along the abscissa. This is a logarithmic
axis, accentuating small CfD values. Finally, several L/h ratios (the horizontal well
length divided by the reservoir thickness, 10, 20, and 30 ft) are plotted. Fig. 5 shows
that for large conductivity fractures (CfD > 20) or small L/h ratios (< 5), the ratio of the
productivity index of a horizontal well and the productivity index of a vertical well

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 11 of 43

[(PI)horizontal /(PI)vertical ] is approximately equal to 1. Therefore, a horizontal well drilled in


the direction of the maximum horizontal stress axis would not be appropriate.
Given a minimum requirement of [(PI)horizontal /(PI)vertical] > 2, a horizontal well drilled so
the hydraulic fracture will propagate longitudinally would be advantageous, if
CfD < 2.5 and L/h = 20.

3.2 Orthogonal Fractures


Hydraulic fractures which propagate orthogonally from the wellbore have two distinct
possible advantages over longitudinal fractures, although their performance is
reduced due to choke skin effect. These are:
1. Multiple parallel fractures are possible.
2. A larger drainage area can be affected.
The location of a horizontal wellbore within the reservoir is the key factor in
determining the benefits to be gained by increasing the potential drainage area. In
some cases, the location of the wellbore within the reservoir may limit the potential of
the well. In the case of orthogonal fractures, the FracNPV* module in the
FracCADE* software can be used to optimize the number and size of hydraulic
fractures. The analysis must incorporate the choke skin effect.
For each orthogonal fracture, the drainage area is divided equally and performance
(including the choke skin effect) over time is calculated. To calculate the skin effect,
some value of the fracture conductivity must be assumed.
The NPV of each fracture is of particular importance to optimize the number of
fractures which are required. It must include, as a fixed cost, a fraction (based on
the number of orthogonal fractures that are planned) of the additional cost of drilling
a horizontal well. The sum of the NPVs must be larger than the NPV of a
hydraulically fractured vertical well. This is illustrated in Fig.6 and Fig.7. Note that
for an individual orthogonal fracture, the optimum fracture length is 800 ft and the
maximum NPV is $1,320,000 compared to 1400 ft and $4,500,000 in the case of a
vertical fracture in a vertical well in the same reservoir.
For this reason, many horizontal wells that are candidates for fracturing require 5 to
10 orthogonal fractures placed along the wellbore to equal the NPV of a massive
hydraulically fractured vertical well. The best candidate reservoirs are usually
relatively thin with low permeability.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 12 of 43

Fig. 6. NPV analysis for a vertical well.

Fig. 7. NPV analysis for one orthogonal fracture.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 13 of 43

3.3 Choke Skin Effect


To demonstrate the impact of the choke skin effect, the expected flow rates from a
vertical well with a vertical fracture and a horizontal well with a single orthogonal
vertical fracture are calculated. Two cases are considered — a reservoir with 1-md
permeability and another with 0.1-md. Well parameters for both cases are listed
below.
h =100 ft
rw =4.872 in.
w =0.25 in.
kf =24,000 md
µ =1 cp
xf =500 ft
p - pwf =1000 psi
xs =10 ft
ws =0.5w
kfs =0.5kf
CfD =10-4
φ=15%
ct =10-5 psi-1
B =1.1 res bbl/STB
The choke skin effect is calculated using Eq. 1. The production is then calculated
and compared after 180 days. The results are shown in Table 1.
Table 1. Performance Comparison Of Vertical And Horizontal
Wells With Fractures
k(md) Vertical Well Vertical Horizontal Well Vertical
Fracture Fracture
1.0 165 BOPD 141 BOPD
0.1 34 BOPD 30 BOPD
As the results clearly show, the reduction in flow rate due to the choke skin effect is
greater when the kh product is larger. However, in a low-permeability reservoir, with
a small kh and small choke skin effect, the production may not justify the additional
cost of drilling the horizontal well section. Therefore, each well must be considered
on a case-by-case basis in terms of NPV and the number of orthogonal fractures
required to maximize economic productivity.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 14 of 43

3.3.1 Fracture Reorientation — Choke Effect


The reorientation of the fracture away from near wellbore stresses may result in
additional choke which may be defined in terms of damaged proppant permeability
or reduction in width of the hydraulic fracture.
πx s k
( S fs )c = (2)
w s k fs
Where:
(Sfs)c = additional choke effect (dimensionless)
xs = length of the damaged interval (ft)
k = permeability (md)
ws = reduced width (ft)
kfs = damaged proppant permeability (md).
To illustrate the potential magnitude and effect of this choke, consider the following
examples using the variables listed above.
For the 1.0-md case, applying Eq. 2, (Sfs)c = 0.125.
For the 0.1-md case, applying Eq. 2, (Sfs)c = 0.0125.
If the hydraulic fracture or fractures will not be propagated either longitudinally or
orthogonally, then it is important to account for the choke skin effect (Sch)c, and a
possible additional choke effect (Sfs)c due to the reorientation of the fracture at some
distance from the wellbore.
The impact of (Sch)c, (Sfs)c and permeability may be illustrated using the values of (Sch)c
and (Sfs)c previously calculated and by assuming that PD(tDxf) = 1.
1
QD = (3)
PD
Where:
QD = rate (dimensionless)
PD = pressure (dimensionless).
1
QD = (4)
PD + ( Sch )c + ( S fs )c
Where:
QD = rate (dimensionless)
PD = pressure (dimensionless)
(Sch)c = choke skin effect (dimensionless)
(Sfs)c = additional choke effect (dimensionless).
Applying , for the 1.0-md case, QD = 0.563 (a 44% reduction in flow capacity). For
the 0.1-md case, QD = 0.928 (a 7% reduction in flow capacity).

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 15 of 43

For hydraulic fractures that propagate orthogonally, the factors that will determine
whether multiple orthogonal fractures will be an economic proposition are the kh
product and the in-situ fracture permeability (kf). In cases when the kh product is
large or the kf is relatively small or both, then the choke skin effect will be large,
resulting in the productivity of each hydraulic fracture being considerably less than
that of a comparable hydraulic fracture in a vertical well.
In the case of a longitudinal hydraulic fracture which propagates within a few
degrees of the axis of the wellbore, the productivity of the fracture will not be reduced
by the effects of either (Sch)c or (Sfs)c. The increased productivity is instead a function
of fracture conductivity and the ratio of the length of the horizontal fracture to the
height of the reservoir. Therefore, longitudinal fractures may be better suited to
relatively thick reservoirs.

3.4 Coning Effects


When considering the possible gains made by hydraulically fracturing a horizontal
well, it is important to not overlook the way that water and gas coning in the reservoir
may be affected by the planned stimulation treatment. These effects may be
modeled using a black-oil reservoir simulator such as “Eclipse” (Ben-Naceur and
Touboul, 1987).

3.4.1 Comparison of Fractured and Nonfractured Reservoir


Assuming the vertical permeability to be equal to the lateral permeability, the
creation of a fracture may significantly delay the occurrence of water coning. Water
breakthrough time is governed by fracture characteristics (length, conductivity and
relative penetration). The value of the vertical permeability has a dramatic effect on
water production, a ratio of vertical to horizontal permeability of less than one-tenth
results in a significant decrease of coning problems.

3.4.2 Effect of the Distance From the Fracture to the Water Zone
Fig. 8 shows the evolution of the water-cut for different vertical penetrations of a
fracture into the water zone for a viscosity (mobility) ratio between oil and water of 2.
For low-mobility ratios, the initial evolution of the watercut versus time is flat, then
increasing linearly with time. For high-mobility ratios, water breakthrough will occur
in a matter of days. Actual breakthrough may occur even earlier due to viscous
fingering.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 16 of 43

Fig. 8. Water breakthrough time versus the position of the fracture


with respect to the oil/water contact.
Production rate (drawdown) also has a dramatic effect on the severity of water coning as shown
in Fig. 9, where the water-cut curves are compared for a constant rate of 100 BOPD and 1000
BOPD.

Fig. 9. Production rate versus water-cut.


Fig. 10 shows the effect of fracture conductivity on water-coning rates for a vertical
permeability of 0.1 kh. Three values of fracture conductivity are considered (10, 1,
and 0.1). After 1000 days, there is a ratio of 2 between the water-cut fractions
corresponding to a 10-fold increase in fracture conductivity.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 17 of 43

Fig. 10. Fracture conductivity versus water-cut.

4 Rock Mechanical Properties


The initiation and propagation of a hydraulic fracture from a horizontal wellbore is a
complex phenomenon. A successful hydraulic fracturing treatment cannot be
designed without understanding the mechanisms that control the initiation and
propagation of a hydraulic fracture. This involves a detailed study of the mechanical
properties of the formation at the borehole and some distance away from it. The
orientation of the borehole with respect to the principle stresses and the relative
magnitude of the stresses are key factors in determining how a fracture initiates and
propagates within the formation.
In cases when fracturing is planned as part of the completion, the mechanical
properties of the reservoir should be determined prior to drilling the horizontal
wellbore section. The borehole can then be oriented, with respect to the principal
stresses, to allow fractures to propagate in the direction that will optimize well
productivity. Fracture orientation versus productivity is discussed in Section 3.

4.1 Openhole Wellbore Stability


The following potential failure mechanisms should to be considered while drilling, or
in the case of an openhole completion.
• shear failure
• tensile failure
• matrix collapse.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 18 of 43

In the case of ductile rock, hydrostatic pressure in the wellbore (less than the radial
stress acting around the wellbore) will result in nonelastic inward movement of the
borehole and in-hole ellipticity.

4.2 Shear Failure


Shear failure is associated with a shearing movement along one or more planes of
failure and is a function of the medium's unconfined compressive strength and the
angle of internal friction. It can result in active inward movement during production
or a passive outward movement during pressurization.

4.3 Tensile Failure


Tensile failure is of the most interest when considering initiation and propagation of a
fracture because the same stresses are present in the formation after running and
cementing a liner.
When the maximum tensile stress near the borehole equals the tensile strength of
the medium, two kinds of tensile failure can be envisioned and the fracture trace can
theoretically be at an angle to the borehole axis. The second kind does not intersect
the borehole, is concentric, and alone is kinematically stable.

4.4 Matrix Collapse


Matrix collapse can occur in a poorly consolidated formation and in rocks which have
unusually large porosity. These formations generally have diagnostic Mohr failure
envelopes (Fig. 11). This failure is associated with a volume reduction and
densification of the medium. It may also be accompanied by a drastic reduction in
permeability and, consequently, a large and sudden decrease in production. Matrix
collapse may be associated with strain hardening, which prevents further
deformation.

Fig. 11. Mohr failure envelope for matrix collapse.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 19 of 43

4.5 Cased-Hole Wellbore Stability


The borehole may be completed in one of the following ways.
• Noncemented slotted or preperforated liner with or without external casing
packers. This is analogous to an openhole situation with the restriction that
progressive failure and squeezing may be reduced due to the presence of the
liner.
• Cemented casing, perforated or slotted locally or along the entire length. This is
a composite configuration consisting of discrete layers; casing, cement sheath
and formation.
Presuming that the casing is not parted” by movement along faults or other tectonic
features such as movement initiated by drilling, fluid loss or leakoff during fracturing,
several modes of failure behind the casing are possible.
Creep
This is progressive, inward, mostly nonrecoverable movement of the formation and
can ultimately cause the casing to fail. Salt, potash and some sands are particularly
susceptible to this time-dependent phenomenon.
Differential Consolidation
As with matrix collapse around an openhole, production may lead to compaction of a
particular horizon, The associated subsidence of the overburden may then separate
the casing by longitudinal tensile-failure.
Perforation Failure
Perforation stability is an important consideration. In general, stress prediction
around a perforation requires numerical simulation. As a first approximation, a
perforation can be regarded as an openhole subjected to a complex loading
condition within the stress field generated by the borehole. Fig. 12 shows an
example of stress contours prevailing at the junction of a perforation with the
borehole. The orientation of the wellbore with respect to the in-situ stress field may
be a stabilizing or destabilizing factor.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 20 of 43

Fig. 12. Stress distribution around a perforation.

4.6 Stress and Deformation Analysis


Plastic effects, nonlinear effects (that is, stress-dependent elastic constants) and
time-dependent effects (that is, viscoelastic or viscoplastic behavior) must be taken
into account if wellbore stability rather than fracture initiation is being considered.

5 Fracture Initiation And Propagation


The fracture initiation pressure, location on the wellbore where the fracture starts,
and initial fracture orientation depend on the stress-state prevailing at the wellbore.
Consequently, the equations developed for borehole stability and openhole elastic
deformation can be used to predict how a fracture will initiate from the wellbore.

5.1 Initiation Pressure


Fracture initiation from the wellbore is a tensile failure process which can be
predicted by the position of in-situ stresses around the wellbore. Tensile failure
occurs when the minimum principal stress exceeds the tensile strength (σo) of a
formation. By pressurization of the wellbore, tensile stresses are induced,
diminishing compressive stress concentration around the wellbore. At the point
where the maximum tensile stress exceeds σo, a hydraulic fracture will initiate. Since
the stress concentration around a wellbore is a function of the wellbore inclination
and orientation with respect to the principal horizontal stresses, the initiation
pressure is also a function of these two variables.
Fig. 13 illustrates this concept. The required initiation pressure is a function of α and
the borehole inclination where α is the angle between the borehole and the minimum
principal stress). The example shows that, as the orientation of the wellbore
changes, so does the pressure required for failure either in tension as the wellbore is

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 21 of 43

pressurized, or in shear as the wellbore is produced. Initiation pressure increases as


the differences between the principal stresses decrease.

Fig. 13. Initiation pressure as a function of α and the borehole inclination.


When the wellbore is drilled parallel to the direction of the minimum horizontal stress
(a = 0), a fracture is initiated at the highest pressure. When α = 90, the fracture
initiates at the lowest pressure.

5.2 Fracture Initiation


In cases when σv < σHmax, fracture initiation will occur at ψ = ο, π where ψ is the angle
measured counterclockwise from the top of the wellbore (Fig. 14).

Fig. 14. Horizontal well configuration in the in-situ stress field.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 22 of 43

At the points of initiation y = 0 and π, the initial fracture will extend at an angle γ
within the tensile region created by the borehole pressure. The angle can be
determined using Eq. 5.
2σ ψz
γ = 1 / 2 tan − 1
σψ − σz (5)

Fig. 15. Initiation points and fracture orientation on the borehole.


Where σψ , σz, and σψ z are the tangential, axial, and shear stresses at the point of
initiation on the wellbore (Fig. 15). σψ , σz, and σψ z are a function of the wellbore
orientation with respect to the principal in-situ stresses (θ); γ is also a function of θ
(Fig. 16).

Fig. 16. Fracture initiation pressure.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 23 of 43

The initial fracture extends in the tensile zone until at some distance from the
wellbore it turns to become normal to σHmin. The maximum length of the initial fracture
can be calculated using Eq. 6.
ψt
L f = D cosec γ sin (6)
2
Where:
Lf = length of the initial fracture (ft)
D = wellbore diameter (ft)
ψt = angle encompassing the tension zone (degrees).
The angle σψ t is determined from the circumferential distribution of tensile stresses
that are induced around the wellbore due to pressurization.
The location of fracture initiation on the wellbore, breakdown pressure and initial
direction of fracture growth along the wellbore can be determined by using the
results of a Differential Strain Curve Analysis (DSCA) performed on an oriented core
sample.

5.3 Fracture Propagation


Hydraulic fracture propagation is very important in terms of possible fracture
interaction in the following cases.
• multiple fractures transverse to the wellbore treated in individual stages
• multiple longitudinal fractures collinear along the wellbore treated in individual
stages.

5.4 Longitudinal Fractures


When the minimum stress is perpendicular to the wellbore and the fractures will not
rotate away from the wellbore, the fractures will not sense the presence of other
fractures until their tips are extremely close. Spacing between the zones to be
treated can be approximately determined on the basis of the desired length of the
individual fractures. Fig. 17 shows the effect of the distance between such fractures
on their width.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 24 of 43

Fig. 17. The effect of distance between collinear fractures on maximum fracture width.

5.5 Angled Fractures


In cases when the wellbore is at an angle to the principal stresses, the fractures will
only initiate collinear to the wellbore and then rotate away from the wellbore.
Normally, inducing a fracture in one zone would alter the existing stress field around
the fracture for a certain distance from it. A second fracture created in this localized
stress field would be subject to these stresses, and hence may not be parallel to the
first fracture. Studies have indicated that the second fracture will rotate toward the
first fracture due to the change of minimum horizontal stress in the localized stress
field. The degree to which a second hydraulic fracture will rotate as a result of the
stress field induced by the first hydraulic fracture is a function of the ratio of minimum
horizontal stress to the maximum horizontal stress and net pressure. Increasing
Young's modulus or propped fracture width results in a higher net pressure and
causes greater rotation of the second hydraulic fracture. The normal distance
between two fractures created from two intervals spaced a distance from each other
in the same wellbore can be determined using Eq. 7.
d = S cosθ (7)
Where:
d = distance between two fractures
S = interval spacing
θ = well deviation from σHmin.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 25 of 43

Once the normalized spacing (2d/h) is greater than 3, little if any rotation takes place
(Fig. 8).

Fig. 18. Fracture rotation angle versus spacing.

5.6 Transverse Fractures


Numerical studies indicate that transverse fractures do not influence each other
unless their dimension, minimum height, or penetration approaches their spacing
(Fig. 19). The interaction entails increased treatment pressures, reduced width and
possible rotation of the fractures (Fig. 20). Interaction is proportional to fracture
width and spacing. A previously propped fracture will influence a propagating
fracture in proportion to its propped width.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 26 of 43

Fig. 19. Width and excess pressure as a function of spacing for parallel,
transverse and radial fractures.

5.7 Controlling Fracture Reorientation


Experiments have shown that when the minimum principle stress is not at an angle
of either 0° or 90° to the wellbore, induced fractures can be nonplanar and S-
shaped.
Two factors control the radius of the curvature, the pump rate and the ratio between
the maximum and minimum horizontal stresses. Analytical and experimental studies
have indicated that the radius of reorientation increases with increasing pump rate,
and that the radius of reorientation decreases as the ratio between the maximum
and minimum horizontal stresses increases (Fig. 20).

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 27 of 43

Fig. 20. Radius of fracture reorientation as a function of the ratio between


the maximum and minimum horizontal stresses.
Qualitatively, an analysis can be made for a wellbore drilled at any angle to the
principal stresses, using stress components normal to, and along the wellbore.
Practically, the ratio between the maximum and minimum horizontal principal
stresses represents a pre-existing state of stress and when it is greater than two, the
radius of reorientation cannot be easily changed by increased pump rate alone.

6 Perforating
Hydraulic fractures frequently propagate in a direction different than the direction
they initiated, due to the relative orientation of the wellbore with respect to the in-situ
stresses.
A fracture will initiate at a preferred location on the wellbore and extend at angle γ
with respect to the borehole (Fig. 15) until at some distance from the wellbore it turns
to become normal to σmin. The initial fracture length is a function of both γ and ψτ.
Consequently, Lf is a function of the angle of the wellbore with respect to the axis of
minimum principal stress. Lf has previously described in Eq. 6. Along the length of
Lf, perforations can be connected to each other by the initial fracture. However,
when two perforations are separated by a distance greater than Lf, two independent
fractures may be initiated. Therefore, Lf may be used as a parameter to estimate the
critical distance between two perforations beyond which more than one initial

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 28 of 43

fracture can be created. Fig. 21 illustrates Lf plotted against the orientation of the
wellbore with respect to the minimum principal stress where D is the diameter of the
wellbore.

Fig. 21. Critical distance between perforation versus well orientation.


In view of the angle γ at which Lf extends along the wellbore, 120° or 60° phased
perforating is recommended at the highest possible shot-density. This will ensure
that the fracture will be in contact with the greatest number of perforations,
minimizing perforation friction pressure and additional choking of the fracture due to
limited contact with the wellbore.
As a general guideline, the distance between the perforations should not exceed the
wellbore diameter and the length of the perforated interval should theoretically not
exceed four times the wellbore diameter, if only one fracture is to be initiated within
the length of the perforated section. The friction pressure across the perforations
may require a longer perforated interval or the use of larger perforating guns. Some
operators use phased guns (0.7-in. perforations, 4 spf) to perforate two to five ft
along the wellbore, commonly three ft.
Penetration is an important consideration in a horizontal wellbore because the casing
will frequently be uncentralized, resulting in a thicker than anticipated cement sheath
on the high side of the wellbore. However, penetration must not be obtained at the
expense of perforation size. A minimum 0.6-in. perforation diameter is required if
the proppant concentration will exceed 6 PPA. The final selection of charges and
guns should be made in conjunction with wireline and testing services.
For an orthogonal fracture, theoretically γ = 90° and the casing should be slotted
rather than perforated to maximize the area of the wellbore in contact with the
fracture. The ABRASIJET* technique is ideal for slotting the casing.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 29 of 43

In the case of a longitudinal fracture, theoretically γ = 0° and, therefore, the length of


the perforated interval should not be critical. However, it is unlikely that a well will be
drilled exactly in the direction of σHmax.
In an extensively perforated or poorly cemented cased-hole completion, unexpected
longitudinal fracture initiation may occur when the well is drilled in the direction of the
minimum stress, due to the influence of the near-wellbore stress field. In this
situation, the fracture will rotate to become perpendicular to the minimum stress
direction as it propagates away from the wellbore.

7 Treatment Design
The principal difference between designing a successful fracturing treatment in a
horizontal well as opposed to a vertical well is understanding the way that the
fracture will propagate from a horizontal wellbore. When fracturing of more than one
interval is planned, care must be taken to avoid possible interference and
communication between the fractures. This requires that a detailed study of the
mechanical properties of the wellbore be conducted prior to designing the fracturing
treatment.
Despite the very obvious differences noted above, the basic steps that have to be
followed when designing a fracturing treatment in a horizontal well are similar to
those for a vertical well. Outlined below are the steps of the procedure required to
design a successful fracturing treatment, taking into account the orientation of the
wellbore with respect to the axis of the minimum principal stress.

7.1 Net Present Value Analysis


An analysis can be used in conjunction with the procedure outlined below to
determine the productivity and NPV of orthogonal fractures in a horizontal well.
However, for a complete analysis, or in the case of longitudinal fractures, it will be
necessary to use a block-matrix reservoir simulator.

7.1.1 Calculating the NPV of Orthogonal Fractures


The following procedure can be used to estimate the NPV from one or more
orthogonal fractures to optimize the number of fractures. The procedure provides an
approximation to transient flow for comparative purposes and is not proposed as a
substitute for a reservoir simulator.
Procedure
1. Select the number of fractures (n) and the distance (L) between them.
2. Determine the drainage radius (re) and the drainage area of each fracture using
an approximate fracture length.
3. Using NPV analysis, optimize the production from the fracture in terms of
fracture geometry.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 30 of 43

4. Using cumulative production versus time, divide the cumulative production by


the number of days the well has produce to determine average production (q)
during this time period. If the NPV program uses a transient solution (rather
than a simplified steady-state solution), this average includes flush production.
5. Calculate the average dimensionless pressure (pD) without a skin using Eq. 8:

∆P 2π kh / µBO
pD = (8)
q
Where:
pD = average pressure (dimensionless)
∆P = P - Pwf (psi)
k = permeability (md)
h = reservoir thickness (ft)
µ = viscosity (cp)
Bo = STB/res bbl
q = time period.
6. Determine the production with a choke skin effect using Eq. 9.

1
qs = α
pD + S (9)
Where:
qS = production with choke skin effect
α = ∆P2πkh/µBO
pD = average pressure (dimensionless)
S = choke skin factor derived from Eq. 1.
7. Determine the ratio of production with and without a choke skin effect (β) using
Eq. 10.
qs
β=
q (10)
Where:
β = ratio of production with and without choke skin effect
qS = production with choke skin effect
q = production without choke skin effect.
Eq. 9 and Eq. 10 can also be used to account for a choke effect due to the
reorientation of the fracture (Sch)c. (Sch)c and (Sfs)c may be combined and
substituted for S.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 31 of 43

8. Determine the total NPV of the fractured horizontal well at a given point in time
using Eq. 11.

NPV fr = n {$ / bbl × discount (Cβ ) − FC } (11)


Where:
NPVfr = total NPV of the fractured well
C = cumulative production without skin (determined using the FracNPV
software)
β = f(qs/q)
FC = fixed cost ($) of one fracture
n = number of fractures.
9. Optimize the number of fractures by repeating this procedure with various values
of n and comparing NPVfr.
The incremental NPV of a fractured horizontal well over an unfractured well can
only be determined of the cumulative production from the unfractured well can
be determined for the same time period. The incremental NPV is determined
using.

NPV Incremental
= NPV fr − NPVh (12)
Where:
NPVfr = total NPV of the fractured well
NPVh = NPV of the unfractured well.

7.1.2 Horizontal Well Production Prediction


The pressure response of a horizontal well will act like the pressure response of a
hydraulically fractured vertical well when the dimensionless length of the horizontal
well is greater than four (Eq. 13).

L kv
LD = (13)
2h kh
Where:
LD = fracture length (dimensionless)
h = fracture height (ft)
kv = vertical permeability (md)
kh = horizontal permeability (md).
The boundary conditions for a fully penetrating infinite-conductivity fracture are
• uniform flux: the same production rate per unit of fracture length

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 32 of 43

• infinite conductivity: no pressure drop along the length of the fracture.


A uniform flux condition is not reasonable to presume in the case of a horizontal well.
Therefore, infinite conductivity is the only valid boundary condition. The following
correlation may be used to calculate the deviation of an infinite-conductivity
horizontal wellbore from that of a fully penetrating, infinite-conductivity fracture of the
same length in a vertical well (Eq. 14).

log S' = A' − B' log LD + C' (log LD ) 2 (14)


This is valid for 0.1 ≤ LD ≤ 100 and 0.125 ≤ zwD ≤ 0.5, and for 0.1 ≤ LD ≤ 25 and
0.0625 ≤ zwD ≤ 0.125. The constants A’, B’ and C’ are given as —
A′ = A1 + A2 log rwD + A3 (log rwD)2
B′ = B1 + B2 log rwD + B3 (log rwD)2
C′ = C1 + C2 log rwD + C3 (log rwD)2
for 10-4 ≤ rwD ≤ 10-2.
The values of A1 through C3 are provided in Table 2.

Table 2. Pseudoskin Factor Correlation Contstants


Well 0.5 0.25 0.125 0.0625
Location
zwD
Constants
A1 -0.8761 -0.5475 0.02620 0.1027
A2 -0.6829 -0.4778 -0.1599 -0.1369
A3 -0.08058 -0.04881 -0.003549 -0.0007541
B1 2.8521 2.4183 2.1247 1.8377
B2 0.9297 0.6929 0.5502 0.4048
B3 0.1243 0.09135 0.07284 0.05342
C1 -1.1258 -1.1817 -1.7673 -1.4129
C2 -0.4764 -0.5726 -0.9499 -0.7704
C3 -0.05145 -0.07205 -0.1296 -0.1067

The correlation shown in Eq. 14 demonstrates that the pseudoskin factor (S′) is
strongly dependent on LD and rwD. S′ (a positive number) should be added to the
horizontal skin factor, Sh (Eq. 15 and Eq. 16).
Sh = 1n ( rw′ / rw ) (15)

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 33 of 43

Sh = S ′ − 1n ( rw′ / rw ) (16)

When S′ is less than 0.5, it indicates that the horizontal well can be treated as a fully
penetrating infinite conductivity vertical fracture. In this case, single-phase
production forecasting may be done as follows.
1. Assuming a square drainage area with dimensions of each side being 2xe,
calculate 2xe /L.
2. Calculate TD for the time of interest (Eq. 17).

0.001055kt
TD =
φµct L2 (17)

3. Determine qD using the values calculated in step 1, step 2 and Fig. 22.

Fig. 22. Single phase flow


The simple solution outlined above is valid for 1 < 2xe /L < 5 and for values of TD
shown in Fig. 22. Beyond this range, an exponential type curve (Fetkovich, 1980)
can be used substituting wellbore radius (rw) for and effective wellbore radius of L/4
for L/xe < 0.4.
In cases when S′ (calculated from Eq. 14) is greater than 0.5, it indicates a deviation
of the horizontal well solution from that of a vertically fractured well. Then a
pseudosteady-state solution can be used to forecast production from the horizontal
well.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 34 of 43

7.2 Fracture Height


The vertical propagation of the fracture versus net pressure must be determined.
This should be done using data from oriented cores taken from the reservoir, as well
as the boundary formations, and can be supplemented by running a FracHite* log in
an offset well. The FracHite log is another way the Mechanical Stability Log data
can be presented. Although the pressure required to initiate the fracture at the
wellbore is dependent on its orientation with respect to the principal in-situ stresses,
the net pressure required to propagate the fracture away from the effect of near
wellbore stresses will be the same as that for a vertical fracture propagating from a
vertical well within the same formation. However, the treating pressure will be
determined by the net pressure required to exceed the closure pressure of the
fracture in the near-wellbore stress field. For this reason, boundaries which are
effective when fracturing offset vertical wells may not be effective in containing a
hydraulic fracture initiated from a horizontal wellbore (Fig. 13).

7.3 Fracture Orientation


The exact orientation of the wellbore with respect to the three principal axes of stress
must be known, as well as the numerical values of these stresses. This will require
that the DSCA technique is used on an oriented core sample, and that the
orientation of the wellbore is known. The orientation of the wellbore with respect to
the axis of principal stress will determine the orientation of the fractures, with respect
to the wellbore and breakdown pressure.

7.4 Fracture Length and Conductivity


When the fractures are orthogonal to the wellbore, fracture length and conductivity
are initially determined in exactly the same way as for a vertical well by using the
FracCADE software. However, the effect of the additional choke skin effect due to
limited contact with the wellbore and a possible pseudoskin factor due to the
changing orientation of the fracture as it propagates, must be taken into account
when considering the NPV.

7.5 Pump Rate


If the wellbore is not aligned with a principal stress axis, then a high pump rate is
required to increase the radius of the fracture as it reorients to align with the in-situ
principal stresses. However, perforation friction pressure must be taken into
account.

7.6 Fracturing Fluid Selection


The fluids used in a vertical well may be used in a horizontal well.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 35 of 43

Limitations
• The fracturing fluid must have proppant suspension properties adequate to
prevent settling in the horizontal wellbore section during injection.
• Crosslinker and breaker schedules should be designed to ensure the fracture
retains near-wellbore conductivity and to ensure that proppant is not dragged out
of the fracture when producing the well. The fracturing fluid viscosity must be
reduced as much as possible prior to flowback.

7.7 Proppant Selection


The proppant type, size and concentration are important parameters for success of
the fracturing treatment.

7.7.1 Mesh Range


The treatment should be tailed-in with larger (or higher-strength) proppant to
maximize near-wellbore conductivity and to offset any effects of irregular fracture
geometry.

7.7.2 Proppant Type


The closure pressure at the wellbore will vary as a function of the wellbore
orientation. When optimizing the proppant selection, this should be accounted for.
Tail-in with a higher-strength proppant should be sufficient. Closure pressure can be
accurately determined using the DataFRAC∗ Service or may be approximated from
DSCA.

7.7.3 Proppant Concentration


The proppant concentration should be ramped-up as high as possible during the
final stages to maximize near-wellbore conductivity and to minimize the pseudoskin
effect. This is more critical when creating multiple transverse fractures.
Tail-in with resin-coated proppants is recommended to prevent proppant flow back
during flowback operations or dropout during production.

8 Execution

8.1 Perforating
Perforating considerations are
• the length of each interval to be perforated

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 36 of 43

• whether to perforate all the intervals individually at the time of the fracturing
treatment or all at one time prior to initiating the fracturing treatments
• perforation density, charge type and perforation size
• orientation of the perforations in the wellbore.

8.1.1 When to Perforate


When to perforate is usually an economic consideration. The possibility of saving
rig-time by perforating all the intervals at one time must not be allowed to outweigh
the operational advantages of perforating each interval at the time of the fracturing
treatment. These operational advantages are
• The packer seat can be pressure tested prior to fracturing each interval.
• Tools activated by annulus pressure can be used in the tool string. This is an
important consideration when the problems associated with mechanical
manipulation of a tool string in a horizontal wellbore.
• Annulus pressure can be applied while fracturing, preventing excessive
differential pressure across the packer.
• When running the tool string into the well, packer elements are not liable to be
damaged by being dragged over perforations.
• The unfractured reservoir will not be exposed to the kill fluid in the annulus for an
extended period of time, with the associated potential problems of lost circulation
and formation damage.

8.2 Wellbore Isolation Between Fractures


A cemented liner is usually necessary to ensure positive zonal isolation. Operational
considerations are:
• the length of interval to perforate
• isolation while fracturing each interval
• fracture spacing to avoid possible interference.
The most common techniques used for wellbore isolation are:
• isolation using mechanical tools
• isolation using proppant plugs
• isolation using viscous plugs.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 37 of 43

8.2.1 Isolation Using Mechanical Tools


In medium- and large radius horizontal wells, the individual fractures can be isolated
with bridge plugs and packers. This method of isolation is expensive and risky. Any
foreign material that falls to the bottom of the wellbore can become a potential
problem. Cleanout procedures must be undertaken with great care before tools are
run into the horizontal section. If the tools become stuck, the process to remove
them and clear the wellbore can be costly and time consuming. Nitrogen services
can be used to improve these cleanout procedures. Stabilized foams may be much
more efficient in removing debris than conventional completion fluids.
The following criteria should be considered when deciding to use, or selecting a tool
string.
• The tool string is only run into the well once. It does not require tripping out of the
wellbore between treatments.
• Fractures are protected from the kill fluid in the annulus even when moving the
tool from one interval to another, or in the event that the tool string must be
“redressed”.
• Circulation or reverse-circulation can be accomplished before and after the
fracturing treatment(s).
• The tool sting is operated with limited pipe movement involving (preferably) no
rotation of the drill string.
• There should be a positive surface indication that the tool string is functioning
properly.
• The diameter of the wellbore after the fracturing treatment should not be
restricted by parts of the tool string left permanently in the wellbore.

8.2.2 Isolation Using Proppant Plugs


Intentional screenout and multidensity/multimesh proppant are techniques used for
wellbore isolation. These two techniques have much less inherent risk and are less
expensive than mechanical tools. A simple cleanout trip to dress-off the end of the
proppant plug should be the only additional workover time. Complete cleanout will
not be as imperative as with bridge plugs and packers because the only tools to be
run along the horizontal section are the perforating tools. Perforating tools have a
larger tolerance between the tool and the casing than the bridge plugs and packers.

8.2.2.1 Intentional Screenout


Another method of isolating the fracture is to attempt to intentionally screen out at
the end of the treatment by pumping ultrahigh proppant concentrations. By design,
this should leave a plug that will ensure isolation. The proppant plug will be most
effective if a curable resin-coated proppant is used. The resin coating can keep the

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 38 of 43

proppant in place and can help prevent proppant settling or movement when the
next set of perforations is pumped into.
The disadvantage of this isolation method is that it is extremely difficult to determine
the proppant concentration that will result in a wellbore screenout. If the well does
not screenout, or if the proppant is not set with a resin to prevent movement, a
channel will form at the top of the wellbore above the proppant and isolation will not
be achieved.

8.2.2.2 Multidensity/Multimesh Proppant


A method of isolation similar to intentional screenout is to follow the last proppant
stage with a multidensity/multimesh proppant stage displaced with a matched
density, high-viscosity fluid. This technique involves cutting the flush volume slightly
short, and allowing fracture closure to occur. The multidensity/multimesh proppant is
then pumped or squeezed into place against the proppant in the fracture. The
multiple densities ensure that some of the material falls to the bottom of the wellbore,
some rises or floats to the top, and other material fills the center portion.
The multidensity/multimesh proppant mixture can be obtained by combining sand
with products used in the INVERTAFRAC* Service. The multiple mesh-sizes provide
a good means of obtaining a pressure differential to ensure fracture isolation. The
various mesh-sizes must be selected carefully to ensure that the smallest material is
larger than the proppant-pack pore-throat openings. This will prevent any further
damage to the proppant-pack permeability. The weighted displacement fluid will
help assure that the material stays together as a plug during pumping operations
and can aid during the compaction process that follows by acting in a method similar
to a wiper plug.

8.2.3 Isolation Using Viscous Plugs


The PROTECTOZONE* service may be used as a method for wellbore isolation.
Using the PROTECTOZONE fluid for isolation will eliminate many of the risks
associated with mechanical tools or proppant plugs. The plug can be easily placed
and does not have the potential inconsistencies of settling or incomplete placement
as does a proppant plug. A breaker is added to the mixture and degrades the plug
to a pumpable viscosity in a predetermined time. This method of isolation does not
require expensive cleanout trips because solids that can interfere with the running of
perforating tools are not introduced into the wellbore.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 39 of 43

8.2.4 Wellbore Isolation in an Openhole Completion


The simplest isolation technique in an openhole completion is to use diverting agents
such as FIXAFRAC∗ J66 Diverting Agent or FIXAFRAC J227 Diverting Agent. This
technique has several disadvantages, the greatest of which is the lack of control over
the spacing of fracture initiation points along the borehole. All of the fractures are
likely to initiate over one small interval along the borehole rather than space
themselves over the entire length.
Another disadvantage is the lack of control of proppant placement near the wellbore
during a diversion stage. Over-displacement of the proppant is likely, leaving an
area of low conductivity in the fracture near the wellbore.

8.3 Flowback
The well should be cleaned-up as quickly and efficiently as possible while minimizing
the amount of proppant produced back from the fracture. This is especially critical
when the fracture has limited contact with the wellbore, and near-wellbore
conductivity is critical to the success of the treatment. Recommendations are
• Prior to flowback, assure that the fracture has closed.
• Flow the well at a reduced rate.
• Ensure that the residual viscosity of the fracturing fluid is minimized by using an
aggressive breaker schedule.
It is unlikely that any proppant that comes out of the fracture during flowback will be
transported to the surface. The natural tendency will be for the proppant to settle on
the low side of the hole. Therefore, the fact that proppant is not observed on the
surface while flowing the well does not mean that the fracture is not producing
proppant.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 40 of 43

9 Overview of the Horizontal Well Treatment-Design Procedure


An overview of important fracture treatment design considerations for horizontal
wells is provided in Table 3.
Table 3. Overview of Fracture Treatment Design Considerations for Horizontal
Wells
Item Consideration
Fracture Orientation DSCA, wellbore inclination and azimuth.
Fracture Height Contained, unconfined — function of net
pressure required to initiate the fracture.
Fracture Model Radial — no boundaries.
PKN — with boundaries.
Fracture Length and Choke skin effect and pseudo-skin due to re-
Conductivity orientation.
Fluid Selection Proppant transport, optimized crosslinker and
breaker schedule.
Proppant Type, size, concentration and strength.
Maximum Proppant Increased concentration at the wellbore.
Concentration
Prepad Optimized using the FracNPV software.
Treatment Volume Optimized using the FracNPV software.
Pad Volume Optimized using the FracNPV software.
Stage Volume/Concentration Optimized using the FracNPV software.
Pump Rate Increased rate to increase radius of re-
orientation.
Perforations Length of perforated interval, phasing density,
friction-pressure.
Tubing Size Friction, tubing movement.
Downhole Tools Movement with sand in the wellbore, weight at
the toolstring.
Surface Equipment Same as a vertical well.
Flowing the Well Back Proppant production — increased choke skin-
effect.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 41 of 43

10 Case History
The well was drilled vertically through the reservoir to determine the magnitude and
orientation of the principal stresses. Oriented cores were obtained from the barriers
and the formation.
Well Data
TVD = 10,048 ft
Casing Size = 9.625, 5/8-in.
Casing Shoe = 7245 ft
Kickoff = 7400 ft
Deviation = 90° at 8000 ft
Openhole = 8.5-in.
Liner Size = 5.5-in, 17 lbm/ft
Liner Top = 6416 ft
Centralizers = 1 every 20 ft
BHST = 150°F (66°C)
Stress measurements were made using a number of different techniques (Table 4),
and the horizontal wellbore section was drilled in the direction of the minimum
horizontal stress.
Table 4. Stress Measurement Techniques
Test Stress Direction Stress Magnitude
DSCA N52-84E σHmin = 8.37 - 11.31 kPa/m
σHmax = 9.5 kPa/m - 12.89 kPa/m
Micro Frac σHmin ≈ σHmax
9.95 kPa/m
100 Mesh Sand σHmin = 24600 kPa ≈ 9.95 kPa/m
Injection σHmax = 25650 kPa ≈ 10.40 kPa/m
Acoustic Anisotropy N 57-80E (Natural
Fractures N85E)
FMS (Natural N80-85E
Fractures)
FMS (After Injection) N55-75E
FracHite σHmin = 11.31 - 13.57 kPa/m
U.S.G.S. E-W

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Horizontal Wells Dowell
Page 42 of 43

The number of required orthogonal fractures and the fracture geometry were
optimized using the FracNPV module in the FracCADE software. Five fractures
would be required to maximize the well performance (Fig. 23).

Fig. 23. NPV analysis of the number of orthogonal fractures.

All the hydraulic fractures were designed as follows:


Gross height = 245 ft
Net height = 25 ft
Fracture half-length = 543 ft
Fracture conductivity = 373 md-ft
Average proppant concentration = 1 lbm/ft
Each interval was fractured separately via 5.5-in. tubing using 128,200 gal of YF140
and 280,000 lbm of 20/40-mesh sand at 40 bbl/min. The average proppant
concentration was 7 PPA. Each treatment was tailed-in with 51,000 lbm of 20/40
mesh resin-coated sand. During the last stage, the proppant concentration was
increased to 17 PPA (4000 lbm total) to maximize near-wellbore conductivity.

DOWELL CONFIDENTIAL
Section 1000
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Horizontal Wells
Page 43 of 43

After the treatment, the well production tested 69 BOPD, 326 BWPD, and 64 Mscfd.
This closely matched the predicted fluid production (Fig. 24).

Fig. 24. Actual versus predicted fluid production.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 1 of 49

APPENDIX A - FRACTURING DESIGN-EXECUTION-


EVALUATION EXAMPLE

1 Well Data................................................................................................................................... 3

2 Reservoir Evaluation ............................................................................................................... 3

3 DataFRAC Service.................................................................................................................. 11

4 Treatment Design................................................................................................................... 22
4.1 Fracturing Fluid Selection .................................................................................................. 22
4.2 Proppant Selection............................................................................................................. 22
4.3 Fracture Length Optimization............................................................................................. 24
4.4 In-Situ Stress Data ............................................................................................................. 27
4.5 Approximate Pumping Schedule........................................................................................ 28
4.6 Placement Design .............................................................................................................. 32
4.7 Production Forecast........................................................................................................... 35

5 Treatment Execution ............................................................................................................. 40

6 Treatment Evaluation ............................................................................................................ 42

7 Post-Fracture Well-Test Analysis......................................................................................... 44

FIGURES

Fig. 1. Review output for all transients. ........................................................................................ 5


Fig. 2. Diagnostic plot - Transient 1. ............................................................................................ 7
Fig. 3. Wellbore storage - Transient 1.......................................................................................... 8
Fig. 4. Horner plot - Transient 1. .................................................................................................. 9
Fig. 5. Multi-rate type curve and derivative match. .................................................................... 10
Fig. 6. Verification plot. .............................................................................................................. 10
Fig. 7. Production decline. ......................................................................................................... 11
Fig. 8. NODAL analysis.............................................................................................................. 12
Fig. 9. DataFRAC job record...................................................................................................... 13
Fig. 10. Step-rate test. ............................................................................................................... 13
Fig. 11. DataFRAC job record (replotted). ................................................................................. 14

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 2 of 49 Example

Fig. 12. Pressure analysis ..........................................................................................................14


Fig. 13. Closure pressure estimation.......................................................................................... 16
Fig. 14. “G” plot. .........................................................................................................................18
Fig. 15. Net Present Value. ........................................................................................................24
Fig. 16. Cumulative production...................................................................................................25
Fig. 17. Fracture height-growth history....................................................................................... 35
Fig. 18. Wellbore fracture width profile. ......................................................................................35
Fig. 19. Fracture height profile. ..................................................................................................36
Fig. 20. Production decline. ........................................................................................................40
Fig. 21. Cumulative production...................................................................................................40
Fig. 22. Post-fracture IPR curves (NODAL analysis)..................................................................41
Fig. 23. Job record. ....................................................................................................................42
Fig. 24. Pressure analysis (Nolte-Smith plot). ............................................................................43
Fig. 25. Job record. ....................................................................................................................44
Fig. 26. Pressure analysis. .........................................................................................................44
Fig. 27. Linear flow regime. ........................................................................................................47
Fig. 28. Vertical fracture, linear flow - Transient 1. .....................................................................48
Fig. 29. Multi-rate type curve and derivative match....................................................................48

TABLES

Table 1. Buildup Test....................................................................................................................5


Table 2. DataFRAC Pressure Record ........................................................................................14
Table 3. Pressure Decline Analysis ............................................................................................18
Table 4. Input Data for FracNPV ................................................................................................25
Table 5. Output From the Inverse Module..................................................................................28
Table 6. Output of the PLACEMENT Simulator..........................................................................33
Table 7. Input Data for MLPP Production-Control Input .............................................................36
Table 8. Input Data for MLPP Production-Control Input .............................................................36
Table 9. Input Data for MLPP Production-Control Input .............................................................37
Table 10. MLPP Output ..............................................................................................................38
Table 11. Fracture Geometry .....................................................................................................44
Table 12. Build-up Pressure Response......................................................................................45

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 3 of 49

1 Well Data
The following is an example of a typical DESIGN-EXECUTE-EVALUATE* technique
for a gas well that will be fracture simulated. The application of the RampGEL*
Service and the CleanFRAC* Service has not been included in this example.
Well Data:
BHST = 210°F (99°C)
Perforations = 10,750 to 10,900 ft TMD
10,161 to 10,316 ft TVD
Tubing = 4-1/2-in.; 15.5 lbm/ft
Casing = 9-5/8-in., 47 lbm/ft to 10,042 ft MD
7 in., 29 lbm/ft to 10,902 ft MD
Net height = 130 ft
Hole diameter = 8-1/2-in.

2 Reservoir Evaluation
Pressure transient analysis is the best method for estimating permeability and skin.
This part of the analysis may be provided by the client. Dowell can provide this
analysis if requested. For complex reservoirs, a regional reservoir engineer or
stimulation specialist should be consulted.
A well test was performed prior to the fracture treatment of this example well. The
well was flowed for 100 hr at 800 Mscf/D and then shut in for another 100 hr. The
pressure buildup response data are provided in Table 1. The test sequence is
shown in Fig. 1.
The known reservoir parameters are:
Total compressibility = 0.000103 1 psi
Gas gravity = 0.65
Viscosity = 0.0247 cp
Porosity = 10%
Net Pay = 130 ft
The data were analyzed using the ZODIAC* (Zoned Dynamic Interpretation, Analysis
and Computation) software from Schlumberger. A commercial well test analysis
program can also be used.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 4 of 49 Example

Fig. 1. Review output for all transients.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 5 of 49

Table 1. Buildup Test


Time Pressure Time Pressure Time Pressure
(hr) (psi) (hr) (psi) (hr) (psi)
0.1008 1651.9 1.608 2903.4 20.16 4735.6
0.1128 1667.2 1.704 2954.8 21.36 4748
0.1224 1679.4 1.8 3004.7 22.08 4754.8
0.1344 1694.6 1.92 3064.9 23.04 4763.2
0.144 1706.7 2.016 3110.9 24 4770.9
0.1536 1718.8 2.16 3176.3 26.4 4788
0.1632 1730.8 2.232 3207.8 28.8 4802.5
0.1728 1742.8 2.304 3238.7 31.2 4815.1
0.18 1751.7 2.16 3176.3 33.6 4826.2
0.192 1766.6 2.4 3278.9 36 4836
0.204 1781.4 2.64 3372.3 38.4 4844.9
0.2112 1790.2 2.88 3458.8 40.8 4853
0.2208 1802 3.12 3538.5 43.2 4860.3
0.2304 1813.7 3.36 3611.8 45.6 4867.1
0.24 1825.4 3.6 3680.2 48 4873.3
0.264 1854.4 3.84 3743.4 50.4 4879.1
0.288 1880.9 4.08 3801.8 52.8 4884.5
0.312 1907.1 4.32 3856.4 55.2 4889.5
0.408 2009.7 4.56 3907.4 57.6 4894.2
0.48 2082.7 4.8 3954.6 60 4898.6
0.552 2150.6 5.04 3998.6 62.4 4902.7
0.576 2172.8 5.76 4114.6 64.8 4906.6
0.624 2216.7 6 4148.3 67.2 4910.3
0.696 2280.4 6.96 4263.5 69.6 4913.8
0.72 2300.3 7.44 4311.5 72 4917.1
0.792 2359.2 8.16 4373.3 74.4 4920.2
0.816 2378.5 8.64 4408.9 76.8 4923.2
0.912 2454.1 9.36 4455.6 79.2 4926
0.96 2490.1 9.84 4482.8 81.6 4928.8
0.984 2507.5 10.8 4529.6 84 4931.3
1.008 2524.7 12 4576.4 86.4 4933.8
1.032 2541.8 13.2 4613.8 88.8 4936.2
1.056 2558.7 14.4 4644.3 91.2 4938.5
1.104 2592.2 15.12 4659.9 93.6 4940.6
1.224 2673.6 16.08 4678.4 96 4942.7
1.344 2748.9 17.04 4694.5 98.4 4944.7
1.44 2806.8 18 4708.8 100.001 4946

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 6 of 49 Example

The data were first plotted on a log-log scale for diagnostics (Fig. 2). Real gas
pseudo pressures are graphed against shut-in time.

Fig. 2. Diagnostic plot - Transient 1.


The data exhibit a unit slope lasting approximately one hour which indicates wellbore
storage. The specialized plot for wellbore storage is a cartesian plot of ∆m(p) versus
∆t, shown in Fig. 3. The wellbore storage coefficient is calculated by the software to
give c = 0.0371 bbl/psi.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 7 of 49

Fig. 3. Wellbore storage - Transient 1.


Infinite-acting radial flow will start at about 1.5 log cycles after the end of wellbore
storage. A more definitive indication is to observe the derivative curve. The
derivative curve during radial flow will approach a constant value and the
dimensionless coordinates equal to 0.5 as shown in Fig. 2. The infinite-acting radial
flow regime in this example appeared at about 80 hr after shut-in. If the test is
terminated earlier than this time, type-curve matching is the only method required for
analysis.
The infinite-acting radial flow is analyzed using the Horner method and is shown in
Fig. 4. From this analysis
permeability (k) = 0.045 md
skin (s) = 3.74
P* = 5060 psi.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 8 of 49 Example

Fig. 4. Horner plot - Transient 1.


Note that the Horner straightline is fitted for the portion of the data shown in radial
flow from Fig. 2.
The data were next analyzed with the type-curve matching method. This method is
particularly applicable to early-time behavior. The log-log plot of ∆m(p) versus ∆t can
be matched with Bourdet, et al type curves to calculate the reservoir parameters
using the ZODIAC software. The solver in the software will generate the type-curve
to match the well test data. The results of the type-curve match are shown in Fig. 5.
A verification plot is shown in Fig. 6. The reservoir parameters from this match are
k =0.054 md
s =5.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 9 of 49

Fig. 5. Multi-rate type curve and derivative match.

Fig. 6. Verification plot.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 10 of 49 Example

The two analyses provide essentially the same result, but the results of the type-
curve match will be used because they utilize the complete range of data.
Additionally, the infinite-acting radial flow requires a longer test period to fully mature.
Production decline curves were then generated to evaluate production before and
after a fracture treatment. The FracNPV* module or MLPP (Multi-Layer Production
Prediction) module from the FracCADE* software can be used for the evaluation.
The FracNPV analysis can provide the direct comparison of fracture and nonfracture
production decline curves and is shown in Fig. 7 for a fracture half-length (xf) of 500
ft and a fracture conductivity (kfw) of 800 md⋅ft. (See Fig. 14 in the Treatment Design
section.)

Fig. 7. Production decline.


Fig. 8 shows the NODAL* production analysis system plot of the producing system.
The post-fracture IPR curve for 50 days is generated using the MLPP module for
comparison with the prefracture IPR.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 11 of 49

Fig. 8. NODAL analysis.


Note that there are some differences of the results of the FracNPV analysis
compared to MLPP analysis. The FracNPV software uses an analytical solution and
the MLPP software uses a semi-analytical reservoir simulator. The MLPP analysis is
more accurate than the FracNPV analysis.

3 DataFRAC Service
A DataFRAC* Service was performed prior to the main propped-fracture treatment.
The first stage of the test was the DataFRAC calibration treatment which involved
injecting 30,000 gal of YF*130HTD into the formation at fracturing rate and pressure.
Fluid injection was then stopped and the pressure decline was monitored as the fluid
was allowed to leak off into the matrix. The injection rate and pressure for this test is
provided in Fig. 9.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 12 of 49 Example

Fig. 9. DataFRAC job record.


The second stage of the test involved pumping 9.5 lbm/gal CaCl2 brine for the step-
rate test. The result of the test is shown in Fig. 10.

Fig. 10. Step-rate test.


NOTE: From step rate test, upper bound of closure pressure = 6600 psi. The
flowback was not monitored for this test.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 13 of 49

Pressure was monitored using a dead-string annulus. The hydrostatic pressure of


the fluid column was 5050 psi. A hydrostatic pressure of 5050 psi was added to the
surface pressure reading to obtain the BHP. The difference between the pressure
prior to shutdown and the initial shut-in pressure was approximately 310 psi. A near-
wellbore pressure drop of 310 psi was subtracted from the data and replotted
(Fig. 11). The data points are shown in Table 2. This data was used for the
DataFRAC analysis.

Fig. 11. DataFRAC job record (replotted).

Fig. 12. Pressure analysis.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 14 of 49 Example

Table 2. DataFRAC Pressure Record


Time Volume Rate BHP Time Volume Rate BHP
(gal) (gal) (bbl/min) (psi) (gal) (gal) (bbl/min) (psi)
7:36:00 0 0 5184 7:57:00 30185 0 6687
7:37:00 8 1.79 5235 7:57:20 30185 0 6678
7:38:00 269 9.07 6373 7:57:40 30185 0 6671
7:39:00 610 7.95 6315 7:58:00 30185 0 6663
7:40:00 1073 16.82 6482 7:58:20 30185 0 6655
7:41:00 2176 38.82 6668 7:58:40 30185 0 6648
7:42:00 4229 50.75 6786 7:59:00 30185 0 6641
7:43:00 6408 52.55 6858 7:59:20 30185 0 6635
7:44:00 8597 52.67 6914 7:59:40 30185 0 6631
7:45:00 10836 51.3 6957 8:00:00 30185 0 6625
7:46:00 12988 50.95 6935 8:00:20 30185 0 6618
7:47:00 15140 50.95 6923 8:00:40 30185 0 6615
7:48:00 17275 52.44 6908 8:01:00 30185 0 6612
7:49:00 19541 52.54 6896 8:01:20 30185 0 6606
7:50:00 21740 52.98 6884 8:01:40 30185 0 6600
7:51:00 23981 52.94 6864 8:02:00 30185 0 6595
7:52:00 26227 53.07 6859 8:02:20 30185 0 6591
7:53:00 28516 53.24 6845 8:02:40 30185 0 6586
7:53:40 29980 53.24 6838 8:03:00 30185 0 6579
7:54:00 30185 0 6838 8:03:20 30185 0 6573
7:54:20 30185 0 6838 8:03:40 30185 0 6568
7:54:40 30185 0 6803 8:04:00 30185 0 6563
7:55:00 30185 0 6775 8:04:20 30185 0 6554
7:55:20 30185 0 6754 8:04:40 30185 0 6550
7:55:40 30185 0 6735 8:05:00 30185 0 6544
7:56:00 30185 0 6722 8:05:20 30185 0 6540
7:56:20 30185 0 6708 8:05:40 30185 0 6534
7:56:40 30185 0 6698 8:06:00 30185 0 6528

The pressure decline was monitored to evaluate the closure pressure and the fluid-
loss coefficient.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 15 of 49

Fig. 13 shows the square root time plot of the pressure decline. A distinct slope
occurs between 6400 and 6600 psi. Taking the intersection of the two straight lines
with different slopes, the closure pressure is estimated to be 6500 psi. This result is
consistent with the upper bound of the closure pressure from the step-rate test.

Fig. 13. Closure pressure estimation.


Fracture Geometry
Fig. 10 shows that during the initial phase of propagation, the pressure increases
with injection which indicates PKN-type behavior. The decreasing pressure after the
initial phase indicates the fracture grows into the bounding layer. A decreasing
pressure profile generally indicates unconfined height growth, keeping in mind that
the pressure will increase again if another barrier is encountered.
The pressure behavior during injection match the P3D model of the PLACEMENT II
module in the FracCADE software (shown in Fig. 12). The P3D model is not
supported by DataFRAC module in the FracCADE software at the present time. The
analysis assumes PKN-type behavior because the net pressure matches the fracture
simulator. Using a radial model provides a very-low net pressure value.
Using the closure pressure of 6500 psi, a pressure decline analysis using the “G”
function plot was performed. The analysis accounts for height growth and tip
recession during shut-in by applying a correction to the standard “G” plot. The P* is
taken from the slope at 3/4 ∆P or mG′ (use “recession” option in the DataFRAC
module).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 16 of 49 Example

The result of this analysis, with the net pressure match, is shown in Table 3, and the
“G” function plot with the corrected slope is shown in Fig. 14. The net pressure
match is obtained by adjusting the Young's modulus (E) and the leakoff height.
The results of DataFRAC analysis are
gross fracture height =160 ft
leakoff height =160 ft
fluid-loss coefficient =0.0044 ft/min1/2
Young's modulus (E) = 3.3 x 106 psi.
DataFRAC Screen

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 17 of 49

Fig. 14. “G” plot.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 18 of 49 Example

Table 3. Pressure Decline Analysis

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 19 of 49

Table 3. Pressure Decline Analysis (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 20 of 49 Example

DataFRAC Pressure Decline Data

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 21 of 49

DataFRAC Pressure Decline Data (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 22 of 49 Example

4 Treatment Design

4.1 Fracturing Fluid Selection


The BHST is 210°F (99°C).
Several fluids are appropriateYF100HTD, YF500HT, and YF600 fluids. A
YF100HTD fluid is selected for this example since the crosslink is reversible and
therefore is much less sensitive to shear history.
YF130HTD is selected for the initial treatment design. The rheological properties
(after two hours at 210°F [99°C]) are
power-law exponent (n′ ) = 0.77
consistency coefficient (K′) = 0.05
Rheology guidelines for YF130HTD fluid are provided in the Fracturing Materials
ManualFluids. Job-specific data was obtained from laboratory testing.
The estimated job time is three hours. After three hours, the fluid viscosity at
170 sec-1 ≈ 440 cp, which is still adequate for proppant transport.
Breaker J481 is used to degrade the fluid. The approximate breaker concentration is
5 lbm/1000 gal.

4.2 Proppant Selection


Fracture conductivity is calculated using Eq. 1.
kf w
C fD = × retained permeability (decimal ) (1)
keff x f

Where:
CfD = fracture conductivity (dimensionless)
kf = in-situ fracture permeability (md)
w = fracture width (ft)
keff = formation effective permeability (md)
xf = fracture half-length (ft).
The required proppant permeability is calculated using Eq. 2.
C fD × k eff × x f
kf = (2)
w × retained permeability
Where:
CfD = fracture conductivity (dimensionless)
kf = in-situ fracture permeability (md)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 23 of 49

w = fracture width (ft)


keff = formation effective permeability (md)
xf = fracture half-length (ft)
w = retained permeability (decimal equivalent).
The permeability of the reservoir is very low (keff = 0.054 md) and a CfD = 10 is used
for the design. A dimensionless fracture conductivity greater than 10 will result in a
pronounced reduction in the effectiveness of production increase.
The estimated retained-permeability factor is 0.2. The retained-permeability factor is
a function of postclosure polymer concentration and breaker concentration. Some of
these data are provided in the Fracturing Materials Manual  Fluids. The
FracCADE software can calculate the retained permeability for borate- and titanate-
crosslinked fluids.
The propped fracture width is approximately 0.1 in. The optimum fracture half-length
is determined using the FracNPV software. The selected fracture length can also be
obtained from the client based on the budget. In this case, the optimum fracture
half-length is 500 ft (see Fig. 15).

Fig. 15. Net Present Value.


The required proppant permeability (using Eq. 2) is
10 × 0.054 × 500
kf =
0.1 / 12 × 0.2
= 162,000 md

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 24 of 49 Example

A proppant permeability of 162,000 md is the minimum required to give CfD = 10.


The closure pressure is greater than 6500 psi (from DataFRAC analysis). Using the
proppant selection guide in Treatment Design, low-density 20/40-mesh intermediate-
strength proppant (ISP) is selected to provide the required proppant-pack
permeability.

4.3 Fracture Length Optimization


The FracNPV module is used to optimize the fracture half-length. The FracNPV
software can also be used to select the optimum fluid, proppant and proppant
concentration.
The plot of Net Present Valve (NPV) versus fracture length is shown in Fig. 15. The
optimum fracture length is 500 to 700 ft. The NPV of a 10-md reservoir is also
shown to demonstrate the importance of reservoir permeability in fracture design.
For the 10-md case, a fracture half-length of less than 200 ft would be sufficient.
The fracture half-length of 500 ft is selected for the design. The determination of
fracture half-length may also be discussed with the client.
Fig. 16 shows the incremental cumulative production for the fracture treatment
generated by the FracNPV software.

Fig. 16. Cumulative production.


The input data used for the FracNPV are provided in Table 4.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 25 of 49

Table 4. Input Data for FracNPV

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 26 of 49 Example

Table 4. Input Data for FracNPV (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 27 of 49

Table 4. Input Data for FracNPV (continued)

4.4 In-Situ Stress Data


The following stress profile was obtained using a long-spaced digital sonic log.
In-situ stress testing was performed to calibrate the stress profile from the log. The
result of the “smooth” calibrated in-situ stress profile is shown below.
Stress Profile

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 28 of 49 Example

4.5 Approximate Pumping Schedule


The INVERSE module in the FracCADE software was used to generate the
approximate pumping schedule. The input fracture half-length of 500 ft was
obtained from the FracNPV analysis.
The output from the Inverse module is provided in Table 5.
Table 5. Output From the Inverse Module

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 29 of 49

Table 5. Output From the Inverse Module (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 30 of 49 Example

Table 5. Output From the Inverse Module (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 31 of 49

Table 5. Output From the Inverse Module (continued)

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 32 of 49 Example

Table 5. Output From the Inverse Module (continued)

4.6 Placement Design


A preliminary pumping schedule for the optimum length and conductivity was
obtained from the INVERSE module of the FracCADE software and validated by
running the schedule in the PLACEMENT module. The schedule was adjusted for a
more practical value. The output of the PLACEMENT simulator is provided in
Table 6. The predicted pressure and fracture geometry are provided in Fig. 17,
Fig. 18, and Fig. 19.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 33 of 49

Table 6. Output of the PLACEMENT Simulator

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 34 of 49 Example

Fig. 17. Fracture height-growth history.

Fig. 18. Wellbore fracture width profile.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 35 of 49

Fig. 19. Fracture height profile.

4.7 Production Forecast


The MLPP (Multi-Layer Production Prediction) module in the FracCADE software is
used for production prediction. The MLPP module was introduced in FracCADE
V3.4. It is a semi-analytical multi-layer reservoir simulator capable of modeling the
performance of commingled multi-layer reservoirs.
The MLPP module uses any one of three production modes
• constant production rate
• constant bottomhole pressure (BHP)
• constant wellhead pressure (WHP).
A constant wellhead pressure of 1625 psi is selected as the production mode. The
constant WHP is the mode of production that is the most common in field practice.
This mode should also be selected for generating the NODAL analysis plot.
Table 7, Table 8, and Table 9 show the input data for MLPP production-control input.
The production mode and tubing flow correlation specification is shown in Table 7.
Table 8 shows the production zone data input for each zone. Fracture geometry
from the PLACEMENT simulator can be imported or entered manually (Table 8).
Table 9 is the multiple-zone data display to show some important data for all
specified production zones. Table 10 shows the MLPP output.
The production decline, cumulative production and post-fracture IPR curves are
provided in Fig. 20, Fig. 21, and Fig. 22.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 36 of 49 Example

Table 7. Input Data for MLPP Production-Control Input

Table 8. Input Data for MLPP Production-Control Input

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 37 of 49

Table 9. Input Data for MLPP Production-Control Input

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 38 of 49 Example

Table 10. MLPP Output

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 39 of 49

Fig. 20. Production decline.

Fig. 21. Cumulative production.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 40 of 49 Example

Fig. 22. Post-fracture IPR curves (NODAL analysis).

5 Treatment Execution
The propped fracture treatment consisted of pumping the fluid in accordance with
the pumping schedule in the PLACEMENT design. The average injection rate during
the treatment was 50 bbl/min. The BHP and flow rate of the job execution is
provided in Fig. 23.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 41 of 49

Fig. 23. Job record.


The final proppant concentration of 13 lbm proppant added was maintained until the
end of the job. The treatment was successfully displaced and placed all 498,000 lbm
of 20/40-mesh ISP into the formation.
The Nolte-Smith plot is shown in Fig. 24. A period of approximately-constant net
pressure occurred at the beginning indicating significant height growth into a stress
barrier. The use of a P3D fracture simulator for the treatment design is appropriate.
The pressure then began to increase slightly, indicating another barrier was
encountered. After about 120 min, the pressure started a more rapid increase. The
(approximately) 150 psi pressure increase is attributed to increased slurry viscosity
as proppant concentration increased, rather than tip screenout indication.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 42 of 49 Example

Fig. 24. Pressure analysis (Nolte-Smith plot).

6 Treatment Evaluation
The evaluation can be applied at several levels.
Execution Evaluation
• Service quality evaluation.
• Evaluation of the pumping of fluids at a scheduled rate and specified
performance.
Post-Job Evaluation
• Evaluation of fracture geometry with the primary evaluation tool being
1. The BHP during the actual treatment.
2. Post-fracture well-test analysis.
• Evaluation of the production results compared to design expectation.
The following evaluations are the post-job evaluation examples based on the BHP
during actual treatment and post-fracture well-test analysis (refer to Treatment
Evaluation for service quality evaluation and Reservoir Evaluation for evaluation of
production results using a well performance tracking form).
The P3D fracture simulator was utilized to evaluate the geometry of the propped
fracture treatment. The “calibrated” set of parameters obtained from the DataFRAC

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 43 of 49

service were used in the propped fracture simulation. Fig. 25 and Fig. 26 show a
very good match using the actual data and the simulator. Table 11 provides the
fracture geometry generated by the simulator using the actual pumping schedule.
The PLACEMENT simulator does not need to be rerun if there is no significant
difference between the actual pumping schedule and the design.

Fig. 25. Job record.

Fig. 26. Pressure analysis.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 44 of 49 Example

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 45 of 49

Table 11. Fracture Geometry

7 Post-Fracture Well-Test Analysis


Post-fracture pressure transient tests are available for this well. The well was flowed
for 100 hr at 4000 Mcf/D after the fracture treatment and then shut-in for another
100 hr. The build-up pressure response is provided in Table 12. The data obtained
from the pre-treatment test (k = 0.054 md) are used for this analysis.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 46 of 49 Example

Table 12. Build-up Pressure Response


Time (hr) P (psi) Time (hr) P (psi)
0.1 4695.5 2.16 4739.8
0.108 4696.1 2.28 4741.2
0.12 4696.9 2.4 4742.4
0.132 4697.7 3.6 4753.6
0.144 4698.5 4.8 4762.7
0.156 4699.2 6 4770.3
0.168 4699.9 7.2 4777.1
0.18 4700.5 8.4 4783.1
0.192 4701.2 9.6 4788.6
0.204 4701.8 10.8 4793.7
0.216 4702.3 12 4798.3
0.228 4702.9 13.2 4802.6
0.24 4703.4 14.4 4806.7
0.36 4708.1 15.6 4810.5
0.48 4711.8 16.8 4814.1
0.6 4715 18 4817.5
0.72 4717.8 19.2 4820.8
0.84 4720.4 20.4 4824
0.96 4722.7 21.6 4827
1.08 4724.9 22.8 4829.8
1.2 4726.9 24 4832.6
1.32 4728.8 36 4856.1
1.44 4730.6 48 4873.3
1.56 4732.3 60 4886.9
1.68 4733.9 72 4898.3
1.8 4735.5 84 4907.7
1.92 4737 96 4915.8
2.04 4738.4 100 4918.1
The data were analyzed using the ZODIAC software. A log-log diagnostic graph of
buildup test data is provided in Fig. 27.
The data exhibit a half slope which is evident from approximately 0.2 hr to 2 hr. This
is a strong indication of the presence of a high-conductivity fracture causing the

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 47 of 49

linear flow pattern. The specialized plot for the linear flow is a cartesian plot of ∆m(p)
versus ∆t1/2 shown in Fig. 28.
The fracture length can be estimated from
kxf2 =11,219 md•ft2
xf = 456 ft.
The inferred fracture length is within 15% of that predicted by the fracture simulator.
The data were next analyzed with the type-curve matching method using the solver
in the ZODIAC software. The results of the type curve match gives CfD ≈ 84.
However, no unique match was obtained. When the CfD, is greater than 30, there is
not enough ∆P in the fracture to show uniqueness.

Fig. 27. Linear flow regime.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Page 48 of 49 Example

Fig. 28. Vertical fracture, linear flow - Transient 1.

DOWELL CONFIDENTIAL
Section 1100
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Appendix A - Fracturing Design-Execution-Evaluation
Dowell
Example Page 49 of 49

Fig. 29. Multi-rate type curve and derivative match.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 1 of 26

APPENDIX B - FRACTURING FLUIDS

1 Introductory Summary............................................................................................................. 3

2 Water-Base Fluids.................................................................................................................... 4
2.1 Polymers .............................................................................................................................. 4
2.1.1 Guar Gum................................................................................................................... 4
2.1.2 Hydroxypropylguar ..................................................................................................... 5
2.1.3 Hydroxyethylcellulose ................................................................................................. 6
2.1.4 Xanthan ...................................................................................................................... 7
2.1.5 Carboxymethylhydroxypropylguar .............................................................................. 8
2.2 Crosslinkers ......................................................................................................................... 8
2.2.1 Borate Crosslinker ...................................................................................................... 8
2.2.2 Organometallic Crosslinkers....................................................................................... 9
2.2.3 Crosslink Rate ............................................................................................................ 9
2.2.3.1 YF100 and YF200 Fluids ............................................................................... 10
2.2.3.2 YF300 and YF400 Fluids ............................................................................... 10
2.2.3.3 YF500 and YF600 Fluids ............................................................................... 10

3 Crosslinked Oil-Base Fluids ................................................................................................. 11


3.1 YF"GO"III and YF"GO"IV Fluids......................................................................................... 12

4 Multiphase Fluids .................................................................................................................. 12


4.1 Foams ................................................................................................................................ 12
4.2 Energized Fluids ................................................................................................................ 12
4.2.1 The Gas Phase......................................................................................................... 13
4.3 Emulsions .......................................................................................................................... 14

5 Acidic Fluids........................................................................................................................... 15

6 Fracturing Fluid Characterization ........................................................................................ 16


6.1 Rheology............................................................................................................................ 16
6.1.1 Shear and Temperature ........................................................................................... 16

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 2 of 26

6.1.2 Shear Rate ................................................................................................................16


6.1.3 Shear Stress .............................................................................................................16
6.1.4 Apparent Viscosity ....................................................................................................16
6.1.5 Newtonian Fluids.......................................................................................................17
6.1.6 Non-Newtonian Fluids...............................................................................................17
6.2 Slurry Rheology ..................................................................................................................19
6.3 Proppant Transport ............................................................................................................21
6.4 Fluid-Loss ...........................................................................................................................22
6.5 Conductivity Damage from Fracturing Fluids .....................................................................23
6.5.1 The Effect of Water-Base Fracturing Fluids on Retained Permeability .....................24

7 Fluid Selection........................................................................................................................26

FIGURES

Fig. 1. The structure of guar. ........................................................................................................5


Fig. 2. The structure of HPG.........................................................................................................6
Fig. 3. The structure of HEC.........................................................................................................7
Fig. 4. The structure of Xanthan. ..................................................................................................7
Fig. 5. Borate crosslinking mechanism.........................................................................................9
Fig. 6. The structure of aluminum phosphate ester chain. .........................................................11
Fig. 7. Power-law exponent of a 40 lbm/1000 gal crosslinked water-base fluid. ........................17
Fig. 8. Consistency coefficient of a 40 lbm/1000 gal crosslinked water-base fluid. ....................18
Fig. 9. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid at 40 sec-1..........18
Fig. 10. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid at 170 sec-1.....19
Fig. 11. The effects of proppant on slurry viscosity of a Newtonian fluid....................................20
Fig. 12. The effects of proppant concentration on friction pressure of a water-base fluid. ........21
Fig. 13. Borehole fluid invasion zones........................................................................................23
Fig. 14. Effects of proppant concentration and porosity on postclosure polymer concentration.25
Fig. 15. Effect of polymer concentration on retained proppant-pack permeability......................26

TABLES

Table 1. Comparison Of Nitrogen And Carbon Dioxide..............................................................13

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 3 of 26

1 Introductory Summary
The fracturing fluid is a critical component of the hydraulic fracturing treatment. The
main functions of the fracturing fluid in a propped fracture are to open the fracture
and to transport propping agent along the length of the fracture. An acid fracture
accomplishes in essence the same goal as a propped fracture, that is, a structure of
much greater conductivity in an otherwise much lower permeability medium.
The rheological characteristics of the fluid are often considered the most important.
However, successful hydraulic fracturing treatments require that the fluids have
some other special characteristics. In addition to exhibiting the proper viscosity in
the fracture, the following characteristics are desirable.
• The fluid should be compatible with rock composition and formation fluids.
• The fluid should exhibit low friction pressure during pumping.
• The fluid should provide good fluid-loss control.
• The fluid should break and clean up rapidly after pumping.
• The fluid should be as economical as possible.
Since the reservoirs to be stimulated vary markedly in terms of temperature,
permeability, rock composition, and pore pressure, many different types of fluids
have been developed to provide the characteristics described above. The following
classes of fluids are available.
• linear water-base fluids
• crosslinked water-base fluids
• crosslinked oil-base fluids
• multiphase fluids (foams, polyemulsions, and energized fluids)
• acidic fluids.
Water-base fluids are used in approximately 70% of all fracturing treatments. Oil-
base fluids account for 5%. Multiphase fluids and acid-base fluids are used in
approximately 25% of all fracturing treatments.
Additives are often added to the fracturing fluid for a variety of reasons. Some of the
more important are — to enhance the viscosity at high temperature, to break the
viscosity at low temperature, and to help control leakoff of the fluid to the formation.
Fracturing fluid components, additives, and additive selection are discussed in
Appendix C Additives.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 4 of 26

2 Water-Base Fluids
Water-base fluids are the most widely used fracturing fluids because of their low
cost, high performance, and ease of handling. Potential problems with water-base
fluids are damage to highly water-sensitive formations and proppant-pack damage
caused by concentrated polymer.

2.1 Polymers
Polymers are water-soluble, high-molecular-weight molecules that can be added to
water to make a viscosified solution capable of suspending propping agents.

2.1.1 Guar Gum


Guar is a long-chain polymer composed of mannose and galactose sugars.
Polymers composed of sugar units are called polysaccharides. The guar polymer
has a very high affinity for water. When the polymer is added to water, guar particles
"swell" and "hydrate," which means the polymer molecules become associated with
many water molecules and unfold and extend out into the solution. The guar
solution on the molecular level can be pictured as long, bloated strands suspended
in water. The strands tend to overlap and hinder motion, which causes an increase
in the viscosity of the solution.
The structure of the guar molecule is usually represented as in Fig. 1. The thought
for many years was that guar consisted of a mannose backbone with galactose side
chains on every other mannose unit (one galactose unit to two mannose units). The
galactose and mannose sugars differ in the orientation of the OH groups on the ring.
Recent studies indicate that the arrangement of galactose units may be more
random, with galactose appearing on two or three consecutive mannose units. Also,
the ratio of mannose to galactose may range from 1.6:1 to 1.8:1 instead of 2:1 as
indicated in Fig. 1.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 5 of 26

Fig. 1. The structure of guar.

2.1.2 Hydroxypropylguar
Guar gum comes from the endosperm of guar beans. The process used to produce
guar powder does not completely separate the guar from other plant materials,
which are not soluble in water. Consequently, as much as 10% of the guar powder
will not dissolve.
Guar can be derivatized with propylene oxide to produce hydroxypropylguar (HPG).
The reaction changes some of the –OH sites to –O –CH2 –CHOH –CH3. The
structure of the HPG molecule is shown in Fig. 2. The additional processing and
washing removes much of the plant material from the polymer, so HPG typically
contains only about 2 to 4% insoluble residue. It has generally been considered to
be less damaging to the formation face and proppant pack than guar, although
recent studies have indicated that guar and HPG cause about the same degree of
pack damage.
Hydroxypropylguar substitution makes HPG more stable at an elevated temperature
than guar; therefore, HPG is better suited for use in high-temperature wells. The
addition of the less-hydrophilic hydroxypropyl substituents also makes the HPG more
soluble in alcohol.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 6 of 26

Fig. 2. The structure of HPG.

2.1.3 Hydroxyethylcellulose
Hydroxyethylcellulose (HEC) is used when a very clean fluid is desired. These
polymers have a backbone composed of glucose sugar units which appears to be
similar to the mannose backbone of guar, but there is a significant difference. Guar
contains hydroxl pairs which are positioned on the same side of the sugar molecule
(cis orientation). In HEC, the –OH groups are on adjacent carbons, but they are on
opposite sides of the ring (trans orientation). The cis arrangement is easily
crosslinked, while the trans is not. Fig. 3 shows the structure of HEC.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 7 of 26

Fig. 3. The structure of HEC.

2.1.4 Xanthan
Xanthan is a biopolymer, produced metabolically by a microorganism. Xanthan
solutions behave as power-law fluids even at very low shear rates, while HPG
solutions become Newtonian. At shear rates less that 10 sec-1, xanthan solutions
suspend proppants better than HPG. Xanthan is more expensive than guar or
cellulose derivatives and is used less frequently. Fig. 4 shows the structure of
xanthan.

Fig. 4. The structure of Xanthan.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 8 of 26

2.1.5 Carboxymethylhydroxypropylguar
Carboxymethylhydroxypropylguar (CMHPG) is a "double derivatized" guar that
contains the hydroxypropyl functionality of HPG as well as a carboxylic acid
substituent. CMHPG is crosslinked with aluminum or zirconium complexes.

2.2 Crosslinkers
Polymers produce viscous solutions at ambient temperature; however, as the
temperature increases, these solutions thin significantly. The polymer concentration
can be increased to offset the thermal effects, but this approach is expensive and
damaging. Instead, crosslinking agents are used to dramatically increase the
effective molecular weight of the polymer by binding polymer chains, resulting in high
fluid viscosities at relatively low polymer concentrations.
Inorganic species such as borate salts and organometallic complexes react with guar
and HPG through the cis-OH pairs. When the polymer solution is concentrated
enough that the molecules overlap (for HPG, at least 0.25% wt/wt), the complex can
react with an overlapping polymer so that the two are linked together. A species is
created that has two times the molecular weight of the polymer alone. Because
each polymer chain contains many cis-hydroxyls, the polymer can be crosslinked at
more than one site. Very high-molecular-weight networks develop, especially under
static conditions, resulting in highly viscous solutions.

2.2.1 Borate Crosslinker


Boric acid and borate salts are used to produce crosslinked fluids with guar and HPG
that are stable to 325°F (163°C). At a pH value greater than 8, an extremely viscous
fluid forms in a matter of seconds. A high pH value is required for crosslinked fluid
stability, with a pH value of 9 to 12 as optimum.
Viscosity control of borate-crosslinked fluids is achieved by adjusting polymer
concentration or crosslinker concentration. Thermal stability of borate-crosslinked
fluids is achieved with delayed activator variance and pH control.
A borate-crosslinked fluid will thin when sheared or heated, but will return to its
original state after shear or heat is removed. The borate crosslink is reversible;
crosslinks form and then break, only to form again.
The borate crosslinking mechanism is shown in Fig. 5.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 9 of 26

Fig. 5. Borate crosslinking mechanism.

2.2.2 Organometallic Crosslinkers


Organometallic crosslinkers were developed for fracturing high-temperature
reservoirs. Titanium and zirconium complexes are used because of their affinity for
reacting with oxygen functionalities (cis-OH), stable +4 oxidation states, and low
toxicity. The upper temperature limit for organometallic-crosslinked fluids is
approximately 350°F (177°C). The stability of the polymer backbone, rather than of
the polymer-metal ion bond is the limiting factor. A well with a BHST greater than
400°F (204°C) can be fractured with these fluids if the treatment is designed to
provide adequate cooldown.
The organometallic-polymer bond is very sensitive to shear. High shear irreversibly
degrades organometallic-crosslinked fluid. Unlike the borate crosslinker, once the
bond between the organometallic crosslinker and polymer is broken, it does not re-
form. Crosslinking occurring in a high-shear region is not desirable because an
irreversible loss of viscosity results.

2.2.3 Crosslink Rate


Crosslinking is a chemical reaction; chemical reaction kinetics apply. The factors
affecting crosslink rate (change in viscosity or molecular weight with time) are:
• fluid temperature
• fluid pH value
• shear conditions
• crosslinker type
• the presence of competing organic molecules (ligands).

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 10 of 26

For example, increasing the temperature or the pH value will usually accelerate the
crosslinking reaction. Fortunately, some of these parameters can be controlled to
slow down the crosslinking reaction so that it does not occur in the high-shear region
(generally 500 to 1,500 sec-1) of the tubulars. Ideally, the crosslinking reaction
should occur in the low-shear region (generally 10 to 200 sec-1) of the fracture.
There are benefits to delaying the crosslink reaction. These are:
• better long-term fluid stability at elevated temperatures
• minimized shear degradation of the fluid
• reduced pipe friction pressure which permits higher injection rates and reduces
horsepower requirements.
Ideally, the crosslink time should be equivalent to the tubing residence time and
viscosity should be building as the fluid leaves the tubulars. This is not likely to
occur, unless there is a rapid and significant temperature change at this point
(physically improbable due to heat transfer and subsequent temperature
equilibrization). Practically, a sort of balancing act may be required, the objective
being to maximize in-situ fluid viscosity as near the wellbore as possible. This will
maintain adequate proppant transport at this critical position in order to minimize
proppant banking. The practical solution may come by allowing a certain degree of
"sacrificial" crosslinking to occur in the tubulars, such that the reaction is proceeding
as the fluid enters the fracture, enhancing proppant transport early near the wellbore,
and accepting loss of some long-term-viscosity potential.

2.2.3.1 YF100 and YF200 Fluids


The YF*100 and YF200 fluids use a borate crosslinker. Crosslink rate is
manipulated by varying the concentration of an activator solution which slowly raises
the fluid pH value.

2.2.3.2 YF300 and YF400 Fluids


The YF300 and YF400 fluids use a titanate crosslinker. Crosslink rate is controlled
by competition for the metal ion between the polymer and a ligand. A competing
ligand is added to effect a delay.
The YF300 and YF400 fluids are low-pH fluids designed to be used as the liquid
phase in carbon dioxide foams or as a base fluid energized with carbon dioxide.

2.2.3.3 YF500 and YF600 Fluids


The YF500 and YF600 fluids may be delayed by using a DUO-VIS system; a dual
crosslinker system in which a fast and a slow crosslinker are used in combination.
The fast crosslinker (a borate crosslinker) ensures that there is adequate viscosity at
the perforations. The slow crosslinker (a zirconate crosslinker), which is accelerated

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 11 of 26

by the heating of the fluid in the fracture, ensures there will be a viscous,
temperature-stable fluid in the fracture.

3 Crosslinked Oil-Base Fluids


Aluminum phosphate ester chemistry is used to viscosify oil for fracturing purposes.
Interactions between the aluminum complexes and phosphate ester molecules
produce a long polymer chain (Fig. 6).

Fig. 6. The structure of aluminum phosphate ester chain.


The structure of aluminum phosphate ester chain. The R groups in Fig. 6 are
hydrocarbon chains that must be soluble in the oil to be gelled. The soluble R
groups keep the aluminum phosphate ester polymer in solution. Generally, the R
groups are hydrocarbon chains containing 1 to 18 carbon atoms. The R groups
have a high affinity for oils such as kerosene and diesel which are comprised of 12-
carbon to 18-carbon (and somewhat higher) chains. Crude oils are composed of a
larger number of different organic compounds and may contain paraffins and
asphaltenes. Some of the high-molecular-weight compounds, especially the
paraffins and asphaltenes, are not compatible with the aluminum phosphate ester
gelling system. Many crude oils can be gelled, but should be tested prior to use.
The viscosity of the standard aluminum phosphate ester fluid is controlled by varying
the quantities of aluminum compound and phosphate ester. High-temperature
performance can be enhanced by increasing the polymer concentration; however,
this can result in very high surface viscosity, making suction out of tanks difficult.
The Fracturing Materials Manual—Fluids provides additional information on
crosslinked oil-base fluids.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 12 of 26

3.1 YF"GO"III and YF"GO"IV Fluids


YF"GO"III can be batch-mixed or continuously-mixed. The temperature range of
YF"GO"III is 100° to 200°F (38° to 93°C).
YF"GO"IV is batch-mixed. The temperature range of YF"GO"IV is 150° to 300°F
(65° to 150°C).

4 Multiphase Fluids
The properties of standard water-base or oil-base fluids can be enhanced by
incorporating a second phase into the fluid. Foams and energized fluids are created
by adding gas (usually nitrogen or carbon dioxide) to water- or oil-base fluids.
Emulsions are created by mixing oil, water or acid, and an emulsifying agent.

4.1 Foams
A foam fracturing fluid is a stable emulsion composed of a liquid (external or
continuous) phase surrounding a gas (internal, dispersed or non-continuous) phase
and a surfactant (foaming agent). The infrastructure of a foam fluid is composed
entirely of bubbles. The liquid phase creates the surface structure of the individual
bubbles. Bubble surfaces contact other bubble surfaces with no free fluid separating
the bubbles. The gas phase is 52 to 96% (vol/vol).
A discussion of foam fracturing fluids is provided in Foam Fracturing.

4.2 Energized Fluids


The primary reason for energizing a fracturing fluid is to eliminate the need for
swabbing or pumping. The infrastructure of an energized fluid matrix is composed of
a liquid with bubbles of gas dispersed throughout. A volume of liquid always
separates the individual bubbles. The gas phase is less than 52% (vol/vol).
The compressed gas functions in two ways.
1. When the pressure is released at the wellhead, expansion towards the wellbore
forces the fluid from the formation.
2. The presence of gas in the liquid reduces the weight of the fluid column so that
normal reservoir drive is sufficient to unload fluids from the well.
Reservoir pressure and permeability influence the fluid flowback after a fracturing
treatment. Flowback in a low-pressure reservoir must result almost entirely from the
effect of compressed gas injected with the fluid. Low reservoir pressure combined
with high permeability allows the gas to expand rapidly into the reservoir. When this
condition exists, localized pressure cannot be maintained long enough to overcome
hydrostatic fluid pressure in the well. The result is ineffective flowback. Long shut-in
times allow the gas to segregate from the liquid and should be avoided.
A localized increase in matrix pore pressure temporarily exists after a fracturing
treatment. The time required for this pressure to dissipate and reach equilibrium

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 13 of 26

with the reservoir will vary from a few minutes to a few days, depending on the
properties of the reservoir and reservoir fluids. This additional pore pressure is
available to help return the fracturing fluid to the surface. The energy in the form of
compressed gas can be used for more effective return of the fracturing fluid.

4.2.1 The Gas Phase


Nitrogen and carbon dioxide are the gases most commonly used in foamed and
energized fluids. Formation characteristics, fluid compatibility, and economics are
major factors that are considered during the decision-making process. Nitrogen is
an inert gas, and is the most frequently used because it is versatile. Carbon dioxide
is more soluble in water than nitrogen so more carbon dioxide is required to saturate
the liquid and create the foam. Carbon dioxide is extremely soluble in oil. Stable-
Oil-Foams normally use nitrogen as the gas phase when the foam quality is greater
then 50%. Table 1 provides a comparison of nitrogen and carbon dioxide.
Table 1. Comparison Of Nitrogen And Carbon Dioxide
Property Nitrogen Carbon Dioxide
Hydrostatic Head Low High
Reactive Inert Yes
Solubility in Water Low Moderate
Solubility in Oil Low High
Surface Tension Reduction None Good
Compressibility High Low
Temperature 100°F (38°C) 20° to 40°F (-7° to 4°C)
In certain applications, carbon dioxide may have certain advantages:
• greater hydrostatic pressure (carbon dioxide is more dense than nitrogen)
resulting in lower treating pressure
• more expansion during flowback (aids in total fluid recovery)
• may prevent or remove water blocks.
Fluids that are saturated with carbon dioxide have low interfacial tension which
reduces capillary pressure and damage. The solubilized portion of carbon dioxide
reduces the interfacial tension of the fracturing fluid to levels as low as those
obtainable by many surfactants. Carbon dioxide has an advantage in that the
carbon dioxide is soluble in the water whereas the surfactant may loose its efficiency
by absorbing onto the rock surfaces.
Carbon dioxide also reduces the viscosity of formation oil. Care must be taken to
ensure that the carbon dioxide and the relatively large proportion of surfactants
pumped into the formation do not create emulsions that could damage the
permeability and reduce productivity.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 14 of 26

Carbon dioxide, unlike nitrogen, is not compatible with all liquid phases. Carbon
dioxide is not recommended in the following fluids.
• YF100, YF100D, YF100HTD, YF200, YF200D, YF500HT, YF600LT, YF600HT,
and YF600UT. Carbon dioxide will interfere with the crosslinking mechanism by
lowering the fluid pH value.
Carbon dioxide tends to accelerate the viscosity development in YF"GO" fluids. At
volumes greater than 30%, carbon dioxide reduces the stability of the fluid and may
prematurely break the fluid.
Carbon dioxide is easily dispersed in the YF300LPH and YF400LPH fluids. The pH
buffer contained in the crosslinker solution maintains a constant fluid pH value of
approximately 4, which simulates a saturated carbon dioxide environment.
Consistent fluid performance is ensured despite variations in the carbon dioxide
concentration or complete loss of carbon dioxide during job execution.
Carbon dioxide is pumped at the wellhead in liquid form. The critical temperature
(triple point) of carbon dioxide is approximately 88°F (31°C). Carbon dioxide is a
supercritical fluid commonly referred to as a gas at temperatures greater than 88°F
(31°C). The transition from liquid to supercritical fluid does not affect the physical
properties of either the carbon dioxide or the foam provided the treating pressure is
greater than 1,080 psi, the critical pressure of the carbon dioxide.

4.3 Emulsions
An emulsion is a dispersion of two immiscible phases such as oil in water, or water in
oil, stabilized with a surfactant. Emulsion fracturing fluids are very viscous solutions
with good proppant transport and fluid-loss properties. The greater the percentage
of internal phase (to a point of inversion), the more resistance there is to droplet
movement, resulting in higher viscosity.
The most common emulsion (Super Sandfrac K-1), termed "polyemulsion," is a
water-external emulsion where viscosified water is the continuous phase and oil is
the discontinuous phase. The fluid is 67% oil and 33% water, stabilized with an
emulsifier. Viscosifying the aqueous phase improves the polyemulsion stability and
significantly reduces friction pressure during pumping. The polymer concentration
used is generally 20 to 40 lbm/1000 gal in the aqueous phase, so the polyemulsion
contains only one-sixth to one-third as much polymer as a typical water-base
fracturing fluid. Since so little polymer is used, a polyemulsion will result in less
conductivity damage (than a water-base fluid) and will clean up rapidly. Viscosity
reduction (break) of the polyemulsion occurs when the emulsifier is absorbed by the
formation rock.
The viscosity and stability of the polyemulsion are dependent on the oil droplet size.
The droplet size decreases as the shear rate increases (rather than shear time), so
the shear rate at the mixing pump must be maximized. This can be accomplished by
operating the pump at the highest speed possible. Recirculating the emulsion
increases the shear time rather than the shear rate and is not very effective.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 15 of 26

In comparison to a water-base fluid the disadvantages of a polyemulsion are:


• higher friction pressure
• higher fluid cost (If diesel is used. If lease crude is used, fluid cost is considered
low)
• significant thinning as the BHST increases.
The Fracturing Materials Manua — Fluids provides additional information on
emulsions.

5 Acidic Fluids
An acidic fluid, usually hydrochloric acid, may be used in carbonate formations as a
fracturing fluid. Portions of the fracture face are dissolved as the acid flows along
the fracture. Since flowing acid tends to etch in a nonuniform manner, conductive
channels are created which usually remain when the fracture closes. The effective
length of the fracture is determined by the following:
• acid volume
• acid reaction rate
• acid fluid loss.
The effectiveness of the acid fracturing treatment is largely determined by the length
of the etched fracture.
While the use of acid as a fracturing fluid eliminates many problems inherent in
propped fracturing, it introduces problems of a different nature. The effective length
of a propped fracture is limited by the distance the propping agent can be
transported in the fracture. Similarly, the effective length of an acidized fracture is
limited by the distance acid travels along the fracture before spending.
The major barrier to effective fracture penetration by acid is excessive fluid loss.
Fluid loss is a greater problem when using acid and is very difficult to control. The
constant erosion of fracture faces while pumping makes deposition of an effective
filter-cake barrier difficult. In addition, acid leakoff is nonuniform and can result in
wormholes and enlargement of natural fractures. This greatly increases the effective
area from which leakoff occurs and makes fluid-loss control extremely difficult.
The Fracturing Materials Manual — Fluids provides additional information on acid
fracturing fluids. A discussion of acid fracturing techniques is provided in Acid
Fracturing.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 16 of 26

6 Fracturing Fluid Characterization

6.1 Rheology
Rheology is the science of the deformation and flow of matter. For fracturing fluids,
the important variable is the apparent viscosity of the fluid as a function of shear,
temperature, and time. These relationships are commonly determined for fluids
(without proppant) in rotational concentric cylinder, capillary, or pipe rheometers.
Very little laboratory rheological testing of fluids containing proppant is done because
of the difficulties in evaluating the rheology of these fluids.

6.1.1 Shear and Temperature


A fracturing fluid experiences wide variations in shear and temperature during a
hydraulic fracturing treatment. High shear is experienced by the fluid during
pumping through the tubulars and perforations. Once in the fracture, the shear on
the fluid is significantly less, but fluid temperature increases until it eventually
reaches formation temperature.

6.1.2 Shear Rate


Shear of fluid in laminar flow can be thought of as a process in which infinitely thin,
parallel planes slide over each other. Shear rate is defined as the velocity difference
between the planes divided by the distance between the planes. The usual rate of
shear reported in viscometric experiments is the value at the wall of the instrument.

6.1.3 Shear Stress


Shear stress is the shearing force per unit area of surface. In most measurements,
the shear stress is determined by measuring the torque exerted on a measurement
bob or by the pressure drop across a tube.

6.1.4 Apparent Viscosity


Apparent viscosity is the shear-stress to shear-rate ratio. The apparent viscosity can
be determined using Eq. 1.
47,880 K ′
µa = (1)
γ 1− n ′
Where:
µa = apparent viscosity (cp)
n' 2
K' = consistency coefficient (lbf-secn ;/ft )
γ = shear rate (sec-1)
n' = power-law exponent (dimensionless).

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 17 of 26

6.1.5 Newtonian Fluids


Newtonian behavior implies that fluids have a constant viscosity at all shear rates.
Water, low-viscosity oils, and gases are examples of fluids that exhibit this behavior.
Fracturing fluids exhibit predominantly non-Newtonian behavior. This means that the
apparent viscosity of the fluid is dependent on the shear that the fluid is experiencing
at the moment of interest. A fracturing fluid may have considerably different
apparent viscosities, depending on the shear that is exerted on the fluid. This non-
Newtonian behavior plays a significant role in the tubular and fracture friction
pressures, and in the proppant-transport capabilities of the fluid.

6.1.6 Non-Newtonian Fluids


The power-law model is used to represent behavior of non-Newtonian fluids. A
straight line is predicted on a log-log plot of shear stress versus shear rate. The
slope of the line is denoted as n' (generally less than one) and is termed the "power-
law exponent." The stress at a shear rate of unity is denoted as K' and is termed the
"consistency coefficient." The n′ and K′ values of fracturing fluids change with
increasing time and temperature; n' tends toward unity and K' decreases.
Fig. 7, Fig. 8, Fig. 9, and Fig. 10 show typical rheology data for a crosslinked water-
base fluid containing polymer at 40 lbm/1000 gal. The data is expressed in terms of
n', K', and apparent viscosities at 40 sec-1 and 170 sec-1. Note the reduction in K' and
viscosity with time and temperature.

Fig. 7. Power-law exponent of a 40 lbm/1000 gal crosslinked water-base fluid.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 18 of 26

Fig. 8. Consistency coefficient of a 40 lbm/1000 gal crosslinked water-base fluid.

Fig. 9. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid at 40 sec-1.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 19 of 26

Fig. 10. Apparent viscosity of a 40 lbm/1000 gal crosslinked water-base fluid


at 170 sec-1.

6.2 Slurry Rheology


Fluids containing proppant account for 20 to 80% of the total volume of a fracturing
treatment, yet very little rheological data exists on these slurries. Determining the
rheological properties of fracturing fluids containing proppant as a function of fluid
composition, flow geometry, temperature, time, and proppant size, density, and
concentration is a considerable problem. The majority of instruments used to
determine rheological properties of fluids without proppant is unusable for these
studies because their geometries will not accommodate the large particles and
concentrations.
Fig. 11 shows the affects of proppant on the slurry viscosity of a Newtonian fluid.
Corresponding data for power-law fluids have not been developed.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 20 of 26

Fig. 11. The effects of proppant on slurry viscosity of a Newtonian fluid.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 21 of 26

The friction pressure of fluid containing proppant will increase as proppant


concentration increases. Fig. 12 shows the effect of proppant concentration on
friction pressures of a water-base fluid containing polymer at 40 lbm/1000 gal. in
2.875 in. tubing. Fig. 12 is for illustrative purposes only.

Fig. 12. The effects of proppant concentration on friction pressure


of a water-base fluid.

6.3 Proppant Transport


The purpose of propping agents in a hydraulic fracturing treatment is to hold the
fracture open and provide a permeable path for fluid flow into the wellbore. The
improvement of well productivity depends on the final propped fracture geometry and
fracture conductivity.
Propped fracture geometry is determined by the settling rates of the proppant in the
fracturing fluid during injection and closure. High proppant-settling velocities during
the treatment may result in the formation of a proppant bank at the bottom of the
fracture. This will increase the risk of proppant bridging, high pumping pressure, and
near-wellbore screenout. Low settling velocities result in more evenly distributed

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 22 of 26

proppant over the total fracture height and the greatest potential for productivity
improvement.
Most fracturing fluids are non-Newtonian with fluid viscosity decreasing as shear
increases. The most important variables affecting proppant settling are the non-
Newtonian characteristics of the fluid, wall effects, and proppant concentration. As
shear rates approach very low or very high values, limiting values of apparent
viscosity are reached. In actual treatments, high shear-limiting viscosity values are
approached in the fracture. However, at the center of the fracture channel, the shear
rate is zero and the fluid viscosity approaches the value for zero-shear viscosity.
Low shear viscosity plays an important role in proppant transport during flow
conditions. The settling velocity equation is a modification of Stoke's law (Eq. 2).
1− n ′
g∆pd p2 g∆pd p2  ut 
ut = +   (2)
18 µ o 18 K ′  d p 

Where:
2
g = gravity acceleration (ft/sec )
∆p = density differential of proppant and fluid (lbm/ft3)
dp = particle diameter (in.)
µo = zero shear viscosity (cp)
K' = consistency coefficient (lbf-secn′/ft2)
ut = terminal settling velocity (ft/sec)
n' = power law exponent (dimensionless).

6.4 Fluid-Loss
Fluid loss to the formation is a filtration process which is controlled by the following
parameters:
• fracturing fluid composition
• flow rate and pressure
• reservoir properties (such as permeability, porosity, pressure and fluid saturation)
• the presence of microfractures, macrofractures, or faults.
Prior to the fracturing treatment, the formation contains a number of fluids
(hydrocarbons, water) with different flow properties. When the fracturing fluid
penetrates the formation, there may be three zones (refer to Fig.13):
• a filter cake with varying thickness (R1)
• a zone invaded by the filtrate (R2)
• a region with the reservoir fluids only (R3).

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 23 of 26

Fig. 13. Borehole fluid invasion zones.


In the first region (R1), polymer and fluid-loss additive particles are deposited and
form a filter cake. The rate of fluid flow through the cake is governed by the Wall-
Building Coefficient (Cw).
The second region (R2) contains the filtrate. The filtrate flow rate is governed by
Darcy's law. Derived from Darcy's law is the Viscosity Control Coefficient (Cv).
The third region (R3) has the flow of the native fluids alone. The compressibility,
viscosity, and relative permeability to those reservoir fluids affect the rate of leakoff
of the fracturing fluid. Fluid loss in this region is governed by the Compressibility
Coefficient (Cc).
A complete discussion of fluid loss is provided in Appendix E — Fluid Loss.

6.5 Conductivity Damage from Fracturing Fluids


Polymer that is concentrated within the proppant pack due to fluid leakoff and
volume reduction during fracture closure is the primary cause of proppant-pack
damage. The polymers used in fracturing fluids are too large to enter the pore
throats of most reservoir rocks and therefore become very concentrated. The
severity of damage increases as the polymer concentration increases and is strongly
dependent on the type of crosslinker. The first point is emphasized by the fact that
foams of either nitrogen or carbon dioxide as the gas phase clean up much better
than typical water-base fracturing fluids. The borate-crosslinked fluids are less
damaging than the organometallic-crosslinked fluids. Breakers will reduce the
severity of proppant-pack damage caused by concentrated polymers. The amount
of retained permeability that can be achieved is directly related to the breaker
concentration; increasing the breaker concentration will increase the fracture
conductivity.

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 24 of 26

Polymer concentration and crosslinker type are even more important than polymer
type when considering the effect of fracturing fluid on proppant-pack conductivity.
During the late 1970s, derivatives of base polymers were developed to make a
"cleaner" polymer. Natural guars were derivatized with propylene oxide to create
HPG. Later, HPG polymers were further derivatized to CMHPG. Recent studies
indicate little or no benefit in using these more-costly polymers based on their
proppant-pack permeability damage.
Optimization of fracture conductivity (kfw) and dimensionless fracture conductivity
(Cfd) can only be accomplished by understanding proppant-pack damage and its
effects on in-situ fracture permeability. The in-situ fracture permeability (kf) has long
been recognized as one of the most limiting factors controlling well productivity. The
in-situ fracture permeability is usually only a fraction of the original clean proppant
permeability value.

6.5.1 The Effect of Water-Base Fracturing Fluids on Retained Permeability


Postclosure polymer concentrations of 200 to 1000 lbm/1000 gal are common and
result in reduced fracture permeability, and therefore fracture conductivity. The
parameters known to affect the degree of proppant-pack damage are:
• postclosure polymer concentration
• initial polymer concentration (surface)
• crosslinker type
• fracturing fluid temperature
• breaker concentration.
Postclosure polymer concentration factors may be approximated (assuming that all
the polymer remains within the proppant pack), using Eq. 3.
ps ( 1 − φ / 100)
p′ = × (3)
cs (φ / 100)
Where:
p' = polymer concentration factor (dimensionless)
ps = absolute proppant density (lbm/gal)
cs = average proppant concentration in fluid (including pad) (PPA)
φ = proppant-pack porosity (%).
The final polymer concentration is then calculated by multiplying the initial polymer
concentration by the polymer concentration factor.
An approximation for polymer concentration factors is represented graphically in
Fig. 14 for a 20/40-mesh Northern White Sand proppant pack. The fracture width is
reduced as closure stress increases, which reduces the pore volume to proppant

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix B - Fracturing Fluids
Page 25 of 26

volume ratio within the fracture, and therefore reduces the proppant-pack porosity.
For example, if 50,000 gal of a 40-lbm/1000 gal borate-crosslinked fracturing fluid
were pumped to place 200,000 lbm of 20/40 proppant, the average proppant
concentration would be 4.0 lbm/gal. Assuming a proppant-pack porosity of 33.5%,
Fig. 14 indicates the polymer concentration factor is approximately 11. Therefore,
the postclosure polymer concentration within the proppant pack would be an average
of approximately 11 times the initial polymer concentration, or 440 lbm/1000 gal.
Polymer concentrations of this magnitude, unless thoroughly degraded, are difficult
to displace, causing significant damage to the proppant-pack permeability.

Fig. 14. Effects of proppant concentration and porosity on postclosure polymer


concentration.
The effect of polymer concentration on the retained proppant-pack permeability is
demonstrated in for linear, borate-crosslinked, and organometallic-crosslinked fluids.
The severity of proppant-pack damage increases as polymer concentration
increases and is strongly affected by the crosslinker type. The linear fluid exhibits
only 12% retained permeability at a polymer concentration of 400 lbm/1000 gal. The
crosslinked fluids impair the retained permeability to an even greater degree than the
noncrosslinked fluid. At the same polymer concentration of 400 lbm/1000 gal, the
borate-crosslinked fluid exhibited less than 5% retained permeability. Fig. 15 also
shows that at any given polymer concentration, the organometallic-crosslinked fluid
is more damaging than the borate-crosslinked fluid. The data presented in Fig. 15

DOWELL CONFIDENTIAL
Section 1200
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix B - Fracturing Fluids Dowell
Page 26 of 26

are considered to be the best case because a proppant concentration of 2 lbm/ft2 is


rarely achieved.

Fig. 15. Effect of polymer concentration on retained proppant-pack permeability.

7 Fluid Selection
The selection of a fracturing fluid involves many compromises imposed by
economical and practical considerations. Guidelines for fracturing fluid selection are
provided in Treatment Design.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 1 of 27

APPENDIX C - ADDITIVES

1 Introductory Summary............................................................................................................. 2

2 Fracturing Fluid Components................................................................................................. 2


2.1 Activators ............................................................................................................................. 3
2.2 Buffers.................................................................................................................................. 3
2.3 Crosslinkers ......................................................................................................................... 3
2.4 Emulsifiers ........................................................................................................................... 3
2.5 Foaming Agents................................................................................................................... 4
2.6 Polymers .............................................................................................................................. 4
2.7 Potassium Chloride .............................................................................................................. 4

3 Fracturing Fluid Additives ...................................................................................................... 4


3.1 Bactericides ......................................................................................................................... 4
3.2 Breakers............................................................................................................................... 5
3.2.1 Breakers for Water-Base Fluids.................................................................................. 6
3.2.1.1 Enzyme Breakers............................................................................................. 6
3.2.1.2 Oxidative Breakers........................................................................................... 6
3.3 Clay Stabilizers .................................................................................................................... 9
3.3.1 Clay Types.................................................................................................................. 9
3.3.2 Clay Control Methods ............................................................................................... 12
3.3.2.1 Ionic Neutralization ........................................................................................ 12
3.3.2.2 Organic Barrier .............................................................................................. 13
3.3.2.3 Particle Fusion ............................................................................................... 13
3.4 Fluid-Loss Additives ........................................................................................................... 13
3.5 Friction Reducers............................................................................................................... 14
3.6 Temperature Stabilizers..................................................................................................... 15
3.7 Surfactants......................................................................................................................... 17
3.7.1 Fluorocarbon Surfactants ......................................................................................... 22
3.7.2 Surfactant Selection.................................................................................................. 22
3.8 Nonemulsifying Agents ...................................................................................................... 23

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 2 of 27

3.8.1 Nonemulsifying Agent Selection................................................................................23

4 Additive Selection ..................................................................................................................25

FIGURES

Fig. 1. Apparent viscosity of 40 lbm/1000 gal crosslinked fluids with methanol and sodium
thiosulfate stabilizers. ......................................................................................................16
Fig. 2. Apparent viscosity of a 50 lbm/1000 gal crosslinked fluid containing sodium thiosulfate
and a 60 lbm/1000 gal crosslinked fluid containing methanol. ........................................16
Fig. 3. Surfactant orientation. .....................................................................................................18
Fig. 4. The wettability of oil/water/rock. ......................................................................................20

TABLES

Table 1. Properties Of Common Dowell Surfactants ..................................................................21


Table 2. The Effects of Wettability Change ................................................................................21
Table 3. Summary of Surfactant Action on Mineral Surfaces .....................................................22
Table 4. Properties Of Common Dowell Nonemulsifying Agents ...............................................24
Table 5. Additive Recommendation Guide .................................................................................25
Table 6. Additive Selection Guide ..............................................................................................26

1 Introductory Summary
A fracturing fluid is not simply a viscosified liquid, such as water and guar polymer or
diesel oil and aluminum phosphate ester polymer. Fracturing fluids are complex
systems containing various additives that are used to modify fluid behavior.
Fracturing fluids commonly contain five or more additives.

2 Fracturing Fluid Components


Certain materials are not considered additives because they are required to make
the base fluid. These are:
• activators
• buffers
• crosslinkers
• emulsifiers
• foaming agents
• polymers
• potassium chloride (KCl).

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 3 of 27

2.1 Activators
Activators are strong bases that enable crosslinking by raising the pH value in the
borate-crosslinked fluids.

2.2 Buffers
Buffers are weak acids or bases that are added to water-base fracturing fluids to
maintain a desired pH value. The buffers will maintain the pH value at a desired
level even if an extraneous acid or base is introduced (for example, through
contaminated water or proppant). This is especially important when using enzyme
breakers. The optimum pH range for enzyme breakers is 3.5 to 5.0. Enzyme
breakers are deactivated when the fluid pH value is greater than 9.0.
Buffers are also used to maintain the proper pH value for crosslinked fluids. This is
important because crosslinking rate and polymer stability are affected by the fluid pH
value. Crosslinked fluids are generally formulated to work best in a narrow pH range
(±0.25 units from the optimum). Guar and HPG can be crosslinked at a pH range of
3 to 10, depending on the type of crosslinker used.
Buffers also promote hydration of the polymer. For example, guar and
hydroxypropylguar (HPG) are treated to be dispersible and nonhydrating at a high
pH value. Initially, the water pH value should be high to allow polymer dispersion.
After the polymer is dispersed, the water pH should be lowered to promote hydration.
The pH value can be lowered by adding an acidic buffer after the polymer is
dispersed.
Combining a slowly-soluble acid with the dry polymer is another approach. The
polymer disperses before the acid can dissolve and lower the pH value.
Acetic acid, adipic acid, formic acid, and fumaric acid are acids used as buffers.
Sodium carbonate and sodium bicarbonate are bases used as buffers. All Dowell
gelling agents contain buffers.

2.3 Crosslinkers
Crosslinking agents enable the individual polymeric molecules to form a complex
network of entangled polymer with the associated hydrated water. This results in
higher molecular weight (higher viscosity) and less freedom of random motion
(greater resistance to deformation) for the solvent and polymer. Not only does
crosslinking result in higher viscosity, it lends stability to viscosity loss with time at
elevated temperatures.
Appendix B — Fracturing Fluids provides additional information for crosslinkers.

2.4 Emulsifiers
Emulsifiers are used to create emulsion-base fracturing fluids. The most common
fluid, Super Sandfrac K-1, is composed of 67% hydrocarbon internal phase and 33%
viscosified water external phase.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 4 of 27

2.5 Foaming Agents


Foaming agents are used to create a stable emulsion composed of a liquid (external
or continuous) phase surrounding a gas (internal, dispersed, or noncontinuous)
phase.

2.6 Polymers
Polymers (gelling agents) are high-molecular-weight molecules that can be added to
a base fluid to make a viscosified solution capable of suspending propping agents.
Appendix B — Fracturing Fluids provides additional polymer information.

2.7 Potassium Chloride


Aqueous solutions of 2% (wt/wt) KCl are routinely used as the base liquid in
fracturing fluids. KCl is added to the water to keep negatively charged clay platelets
in place by surrounding them with an electrically neutral fluid containing positive ions.
KCl is used more often than sodium chloride (NaCl) or ammonium chloride (NH4Cl)
because the single charge density and small size of the potassium ion better
stabilizes clays against invasion of water and, consequently, prevents swelling. KCl
helps maintain the chemical environment of the clay particles, but it does not provide
permanent stabilization.

3 Fracturing Fluid Additives


Fracturing treatment optimization involves selecting the best base fluid system
together with special additives required for the specific reservoir and well. Additives
include:
• bactericides
• breakers
• clay stabilizers
• fluid-loss additives
• friction reducers
• temperature stabilizers
• surfactants
• nonemulsifying agents.

3.1 Bactericides
The polymers used in water-base fracturing fluids are excellent food sources for
bacteria. Bactericides are added to water-base fracturing fluids to prevent bacterial
degradation of the polymer and to protect the formation from bacterial growth.
Common practice is to add a bactericide to the frac tanks before water is added to

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 5 of 27

ensure that the bacterial enzyme count remains low. No water-base fracturing fluid
should be pumped into a well without some type of bactericide present. Bactericides
are unnecessary in oil-base fracturing fluids.
Bacteria may be aerobic or anaerobic. Aerobic bacteria require oxygen for survival.
Anaerobic bacteria can exist in the absence of oxygen.
Aerobic Bacteria
Aerobic bacteria produce enzymes that degrade the polymers used in water-base
fracturing fluids resulting in a viscosity loss and premature break of the fluid.
Anaerobic Bacteria
Anaerobic bacteria, introduced by the fracturing fluid, can create severe reservoir
problems. The bacteria can multiply in such numbers that they reduce permeability
and, consequently, damage the formation. Certain types of anaerobic bacteria
chemically reduce sulfate ions to produce hydrogen sulfide, creating a safety hazard.
Hydrogen sulfide also corrodes tubular goods and production equipment.
Bactericide M76
Bactericide M76 is a quaternary amine that will kill bacteria and deactivate the
enzyme. M76 concentrations are dependent on fluid surface temperature and range
from 0.5 gal/1000 gal to 1.5 gal/1000 gal.
Dryocide
Dryocide will kill bacteria and deactivate the enzyme. Dryocide is used at a
concentration of 0.5 lbm/1000 gal.
Microbiocide M275
Microbiocide M275 is a concentrated isothiazolin compound that is adsorbed onto an
inert solid for ease of handling. M275 will kill bacteria but will not effectively
deactivate the enzyme. M275 concentrations are dependent on fluid surface
temperature and range from 0.3 lbm/1000 gal to 0.6 lbm/1000 gal.

3.2 Breakers
Thermal breaking of the polymer backbone generally occurs in wells with bottomhole
temperatures greater than 225°F (107°C). A breaker should be added to the
fracturing fluid when the bottomhole temperature is less than 225°F (107°C).
Breakers are added to fracturing fluids for two reasons.
1. To reduce the viscosity of the fluid so that the fracturing fluid can be cleaned up
quickly following a treatment.
2. To degrade the fluid and thus reduce proppant-pack conductivity damage.
Breakers perform both these tasks by attacking the backbone of the polymer and
reducing its size. The fluid viscosity decreases as the polymer molecular weight
decreases.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 6 of 27

Controlling the timing of the breaking process is critical to the success of the
fracturing treatment. Once a breaker is added to a fracturing fluid, the degradation
process begins immediately. Careful planning must go into the design of breaker
schedules. If too much breaker is added early in the treatment, the viscosity
required for fluid-loss control and proppant transport may be prematurely lost,
resulting in a screenout. If the breaker schedule is not ambitious enough, the
molecular chains may not sufficiently degrade, causing the treatment to clean up
slowly. Even worse, without sufficient breaker quantities, the fluid may not
completely degrade, limiting well production because of proppant-pack damage.
The postclosure polymer concentration in the fracture may increase to
concentrations greater than 500 lbm/1000 gal because of filtrate loss. The breaker
level in this concentrated fracturing fluid decreases as the polymer concentration
increases. The breaker is dissolved into the water portion of the slurry and is lost in
the fluid leaking off as the fracturing fluid dehydrates. This can result in a damage to
the proppant pack which may exceed 90%.
Tapered breaker schedules allow much greater quantities of breaker to be added to
the fluid while minimizing the risk of excessive degradation. To design the breaker
schedule, the time of exposure to bottomhole temperature for each stage of the
fracture treatment must be determined. From this, a maximum quantity of breaker
can be calculated without risking premature loss of viscosity. Detailed breaker
schedule design information is provided in Treatment Design.

3.2.1 Breakers for Water-Base Fluids


The breakers currently used in water-base fluids are enzyme breakers and oxidative
breakers.

3.2.1.1 Enzyme Breakers


Enzyme breakers such as hemicellulase begin to degrade the polymer immediately.
These enzymes are similar to those that bacteria use to digest the polymer.
Enzyme Breaker J134L
Enzyme Breaker J134L is a low-temperature breaker and is the preferred breaker for
low-pH fluids such as those foamed with CO2.
The optimum temperature range for J134L is 75° to 130°F (24° to 54°C). It is not
effective at temperatures greater than 150°F (66°C).
The optimum pH range for J134L is 3.5 to 5.0. In fluids with a pH value greater than
9.0, J134L becomes inactive and should not be used.

3.2.1.2 Oxidative Breakers


The most common oxidative breakers are peroxydisulfates. Thermal decomposition
of peroxydisulfate produces highly reactive sulfate radicals which attack the polymer
backbone. Oxidative breakers are used in applications where the fluid is exposed to

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 7 of 27

bottomhole temperatures greater than 125°F (52°C). Sulfate radical generation


occurs rapidly at temperatures greater than 125°F (52°C). Thermal decomposition is
slow at temperatures less than 125°F (52°C). These breakers can be expanded into
lower temperature applications (60° to 125°F [16° to 52°C]) if an amine is
concurrently added to catalyze the reaction.
Oxidative breakers are effective over a wide pH range (3 to 14) and demonstrate
breaking properties superior to enzyme breakers (based on observed proppant-pack
permeability reduction). This is especially true when the oxidative breaker reaction is
catalyzed by the amine.
Breaker J218
J218 is an ammonium persulfate oxidative breaker. J218 is effective in the
temperature range of 125° to 225°F (52° to 107°C).
In high-temperature wells, J218 can be used in the latter stages to promote rapid
degradation of the polymer at the wellbore or near-wellbore area. This is extremely
useful where forced-closure techniques are practiced.
EB-Clean J475 Breaker
EB-Clean* J475 Breaker is used as component of the CleanFRAC* Service. J475 is
a 20/40-mesh material produced by coating (encapsulating) an ammonium
persulfate oxidative breaker material (J218) with a water-resistant barrier.
Encapsulation of the breaker greatly reduces fracturing fluid exposure to the breaker
and enables the use of high concentrations of breaker that without coating, would
rapidly reduce the fluid viscosity. Unlike breakers that dissolve in the fracturing fluid
(for example, J218 and J134L), J475 will not leak off and be lost to the formation. It
remains in the fracture to degrade concentrated polymers. After the fracturing
treatment, release of the breaker occurs as reservoir temperature increases and the
fracture closes.
An upper temperature limit of 180°F (82°C) is supported by laboratory data. J475 is
most effective when fluid temperatures are predicted to be less than 180°F (82°C).
J475 may rapidly deactivate when fluid temperatures are greater than 180°F (82°C)
resulting in minimal conductivity improvement.
Field data has indicated positive results using J475 at temperatures greater than
225°F (107°C). Because an upper temperature limit for J475 has not been
established during field use, J475 may possibly be used at higher temperatures if
positive results have been observed on other fracturing treatments. Some of these
results are:
• returned fluid had a low viscosity and viscosity did not increase
• cleanup was at least as good as expected
• production was as good or better than expected.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 8 of 27

EB-Clean LT J479 Breaker


J479 is used as part of the CleanFRAC* Service. J479 is a 20/40-mesh material
produced by coating (encapsulating) breaker material (J218) with a water-resistant
barrier. Encapsulation of the breaker greatly reduces fracturing fluid exposure to the
breaker and enables the use of high concentrations of breaker that without coating,
would rapidly reduce the fluid viscosity. J479 is used in low-temperature wells where
J475 is often not suitable. Release of the breaker occurs gradually with time and is
enhanced by low levels of closure stress. J479 is used in conjunction with J318 or
J466. J479 is used when fluid temperatures are predicted to be less than 125°F
(52°C).
Breaker J481
J481 is an oxidative breaker composed of sodium bromate. J481 is approved only
for use with YF*100HTD fluids. The reactiveness (thermal decomposition) of J481
and YF100HTD fluids is strongly dependent on temperature. J481 is effective at
temperatures ranging from 200° to 275°F (93° to 135°C).
Liquid Breaker Aid J318
Liquid Breaker Aid J318 is used in conjunction with oxidative breakers when the fluid
temperature is less than 125°F (52°C). J318 may be used in the WF100, WF200,
YF100 (nondelayed), and YF200 (nondelayed) fluids. The amines in J318
accelerate the generation of sulfate radicals from the oxidative breakers, making it
an effective breaker at a low temperature. J318 is a catalyst and not a breaker on its
own. The application of the amine breaker aid with the persulfate breaker is
patented Dowell technology (U.S. Patent No. 4,250,044, and No. 4,560,486).
Breaker Aid J466
Breaker Aid J466 is used in conjunction with oxidative breakers when the fluid
temperature is less than 125°F (52°C). J466 is used in the YF100D and YF200D
fluids only. (J318 is not compatible with the YF100D and YF200D fluids. These
fluids use a mechanism that slowly raises the fluid pH value and enables delayed
crosslinking. J318 is strongly alkaline and will rapidly raise the fluid pH value.)
Breakers for Crosslinked Oil-Base Fluids
Breakers for crosslinked oil-base fluids (gelled oils) operate much differently than
their water-base counterparts. Most gelled-oil breakers are slowly soluble bases
such as lime or bicarbonate and are intended to reverse the reactions between the
aluminum complexes and phosphate ester molecules. Gelled oils can be difficult to
break at bottomhole temperatures less than 100°F (38°C). The breakers used in
gelled oils are:
• FIXAFRAC*J59
• Breaker J603

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 9 of 27

• Soda Ash M3.

3.3 Clay Stabilizers


Clays are layered particles of aluminum and silicon oxide that are less than four
microns in diameter. Clays can swell or migrate (or both), resulting in permeability
damage. The damage can occur during drilling, completion, or production of the
well. Severity of damage caused by clay swelling and clay migration is dependent
on the following parameters:
• clay type
• clay content
• clay distribution
• pore-size distribution
• grain-size distribution
• quantity of cementing materials (for example, calcite, siderite, and silica).
Susceptibility to damage is evaluated using X-ray diffraction, a scanning electron
microscope, and thin-section point counting.
Clay Swelling
Clay swelling is caused by the introduction of incompatible fluid or relatively fresh
water (water of lower salinity or ionicity than the original pore solution) into the pores.
The expansion of the crystal unit (interlayer distance) is caused by the replacement
of ions which neutralize the charge deficit by dilution and osmotic effect. Therefore,
water or large dissolved ions, that are incorporated in the structure, increase the
distance between layers to compensate for the electrostatic repulsion due to the net
negative charge of each layer. The expansion of each crystal unit results in a
macroscopic swelling of the clay particle. The swelling can seal the pore throats,
resulting in permeability reduction.
Clay Migration
Negatively charged particles result when the charge balance between positive
(aluminum) and negative (oxygen) is disrupted through displacement of cations or
breaking of the particles. Cations from solution surround the clay particle, creating a
positively charged cloud. Such particles repel each other and are prone to migration.
Once clay particles are dispersed, they can block pore spaces in the rock and
reduce permeability.

3.3.1 Clay Types


Kaolinite, illite, chlorite, smectite, and mixed-layer clays are the most common clay
types. Kaolinite, Illite and chlorite are migratory clays. Smectite is a swelling clay.
Mixed-layer clays can swell or migrate (or both).

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 10 of 27

Clays exhibit a cation exchange capacity. The ionic atmosphere and stability of the
clay are a function of the clay type, the cations of the surface of the clay, and the
surrounding fluid.
Clays in the sodium state disperse/migrate when exposed to fresh water. This
occurs because of the rapid diffusion of the sodium ions into the water, creating a
strong repulsion between clay particles.
Kaolinite
Kaolinite is a migratory clay that is found in most sandstones. It is a 1:1 clay
consisting of one tetrahedral (silica) sheet and one octahedral (alumina) sheet. The
different layers are bound together because of the proximity of the hydroxyl ions of
the tetrahedral sheet and the oxygen ions of the octahedral sheet. The hydrogen of
the hydroxyl groups is bound at the same time with oxygens of the octahedral and
tetrahedral sheets. The bond is very rigid. There is no substitution by other ions of
either the aluminum in the octahedral sheet or of the silicon in the tetrahedral sheet.
Therefore, there are no charge deficits on the faces of kaolinite crystals.
In many sandstones, kaolinite is characterized by its loose attachment to the host
grains and the large size of the individual crystals. The crystals are usually bound
together in compact aggregates and are too large to be transported by moving fluids
in the pore system. If these aggregates disperse, fine crystals will be liberated and
entrained in the moving fluid. This is particularly true in areas of high fluid turbulence
(for example, close to the wellbore). These migrating kaolinite crystals can go to a
pore throat where they will lodge and act as a check valve. Thus, the migration of
kaolinite depends on the state of dispersion or aggregation (flocculation) of the
individual crystals.
Illite
Illite is a 2:1 migratory clay consisting of two tetrahedral (silica) sheets and one
octahedral (alumina) sheet. The octahedral sheet is between the tetrahedral sheets.
The oxygens at the tip of the tetrahedra point toward the center octahedral sheet and
substitute for two-thirds of the octahedrally coordinated hydroxyls.
The most common illite mineral has approximately one-half of the silicon substituted
by aluminum in the tetrahedral sheet. Substitutions in the octahedral sheet are low;
approximately three-fourths of the octahedral ions is aluminum, a minor amount of
ferric iron is present, and approximately one-eighth if the cations is divalent
(magnesium and ferrous iron). This gives a total negative charge of approximately
0.75 which is due to substitution in the tetrahedral sheet. Because these
substitutions are balanced by interlayer cations (potassium) close to this sheet, there
is a strong bond between them and, therefore, there is a strong bond between the
different layers. Because of this, the potassium ion in illite cannot be easily removed
and replaced by ions or water molecules (or both). For this reason, illite is
considered nonswelling.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 11 of 27

Chlorite
Chlorite is a migratory clay consisting if 2:1 type layers bound by a magnesium
containing octahedral sheet. The bond between layers is strongly ionic, like illite.
Chlorite is a nonswelling clay. Chlorite contains iron (Fe+3 and Fe+2) substituted for
aluminum, silicon, and magnesium.
Chlorite is readily attacked by hydrochloric acid (HCl) and the iron liberated during
dissolution can precipitate as a gelatinous ferric hydroxide when the acid spends.
The ratio of ferric to ferrous iron in chlorite is typically 0.2. Therefore, iron problems,
if any, can be avoided by using reducing or chelating agents (or both) in the acid.
The shape of the chlorite mineral is similar to a "honeycomb" structure. This creates
a microporosity which, like illite, allows water to be trapped and retained.
Smectite
Smectite is a swelling clay. Normally, migratory clays are considered to have greater
damage potential. Smectite can create severe plugging if it swells in the pore-throat
lining.
Smectite is characterized by the loosely bound cations and layers of water (or polar
organic molecules) between the silica sheet. These are loosely bound because
smectite has most of its charge originating in the octahedral sheet. The resulting
charge deficit in the layer is balanced by an interlayer cation separated from the
octahedral sheet by the tetrahedral sheet. Thus, the interlayer width is reversibly
variable. The interlayer water can be driven off at temperatures greater than 250°F
(121°C). Sodium, calcium, hydrogen, magnesium, iron, and aluminum are the
interlayer cations commonly found.
The introduction of incompatible or relatively fresh water (water of lower salinity or
ionicity than the original pore solution) into the pores is the cause of smectite
swelling. Smectites can be found in flocculated aggregates which can disperse,
leading to the migration of the smectite particles.
Mixed-Layer
A large number of clays are not pure minerals, but consist of interstratified units of
different chemical compositions. These clays are called mixed-layer clays and can
either migrate or swell.
Illite-smectite (illite-montmorillonite) and chlorite-smectite are common mixed-layer
clays with illite-montmorillonite being the most abundant. The two layers occur in all
possible proportions ranging from 9:1 to 1:9. Many of those with a 9:1 or even 8:2
ratio are called illites or glauconites (all glauconites have some interlayered
montmorillonite). Those having ratios of 1:9 and 2:8 are called smectite
(montmorillonite).
Glauconite is also used as a rock name and is applied to any aggregate of fine-
grained, green-layer minerals. The iron illite layers commonly occur interlayered with

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 12 of 27

montmorillonite-like layers. In glauconites, more than one-half of the octahedral


positions is predominantly filled with ferric iron.

3.3.2 Clay Control Methods


Three methods of clay control are known to be effective. These are:
1. Ionic neutralization
2. Organic barrier
3. Particle fusion.

3.3.2.1 Ionic Neutralization


Ionic neutralization is commonly used in hydraulic fracturing and is accomplished
using brines, polyvalent inorganic cations, and quaternary amine polymers.
Brines
In the formation, the clays are generally not dispersed as long as their chemical
environment is not changed. For this reason, brines are not nearly so damaging to
sandstone as is freshwater. Aqueous solutions containing 1 to 3% (wt/wt) salt are
normally used as the base liquid in fracturing fluids. KCl is used more often than
NaCl or NH4Cl because K+ stabilizes clays better against invasion of water and,
consequently, prevents swelling. All of these salts help maintain the chemical
environment of the clay particles, but they do not provide permanent stabilization.
Clay Stabilizer L237 is a temporary clay stabilizer but is more effective than KCl,
NaCl, or NH4Cl. L237 will prevent formation damage caused by dispersion or
swelling of clays induced by any following fresh water. Therefore, L237
concentrations can be reduced in the proppant-laden fluid stages.
Inorganic Polynuclear Cations
The attraction between a negatively charged clay particle and its exchangeable
cations is exponentially related to the charge on the cations. Thus, a polynuclear ion
with a net charge of +8, +12, or more may be several million times more attracted to
a clay particle than monovalent or divalent cations. Consequently, from electrostatic
considerations alone, polynuclear ions should almost immediately displace all of the
exchangeable cations and be very tightly held to the clay surface.
Clay Stabilizer L42 is the inorganic polynuclear cation used by Dowell. Zirconium
oxychloride is hydrolyzable metal ion forming a polynuclear cation with a high
cationic charge. L42 is a permanent clay stabilizer used primarily in acidizing
applications.
L42 will crosslink most fracturing fluids and, typically, is not used in fracturing
operations.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 13 of 27

Amine Polymers
A monomolecular film of quaternary amine polymer is strongly adsorbed on the
surface of the clays by cation exchange. More permanent stabilization is claimed
since the clay particles are bridged together by multiple cationic sites along the
polymer chain. To destabilize the clays, simultaneous release of all cationic sites is
required for exchange with other ions in the formation brine. Quaternary amine
polymers are water soluble and leave the formation water-wet. Quaternary amine
polymers may be used in water-base fracturing fluids under acidic, neutral, and basic
conditions.
Clay Stabilizer L55 is the quaternary amine polymer used by Dowell and is normally
used in reservoirs with clay content greater than 10%.

3.3.2.2 Organic Barrier


Some cationic surfactants prevent deflocculation of clays by their adsorption on the
clay surface (cation exchange). These cationic surfactants make clay and
sandstone surfaces oil-wet. The oil-wet condition prevents the adsorption of water
which would otherwise deflocculate the clays. For adequate stabilization, greater
concentrations of cationic surfactants may be required as anticipated by cation
exchange due to physical adsorption of a second layer of surfactant, where the new
surfactant molecule is held to the former one by hydrocarbon bonding through their
hydrophobic tail. This method temporarily reverses the wettability of the rock,
resulting in a decrease in the permeability of the oil due to water entrapment. In
most cases, reversing the wettability of the rock is undesirable.
The organic barrier method is not used by Dowell.

3.3.2.3 Particle Fusion


This method is technically an anionic stabilization. In this method, the repulsive
forces between the dispersing clay particles are destroyed by simply destroying part
of the clay mineral itself, principally the octahedral sheet. This can be accomplished
by the use of the various anions of fluoride, phosphate, borate, and to some extent
hydroxide. When these anions attack the clay platelets, the aluminum from the
octahedral layer and the components of the injected chemical compound combine to
form a very thin connecting sheet over the remaining partial and the complete
(unaltered) clay platelets. This connecting sheet binds all of these formerly
dispersible clay particles together.
The particle fusion method is not used by Dowell.

3.4 Fluid-Loss Additives


Fluid-loss control is essential for an efficient and successful fracturing treatment.
The loss of fracturing fluid into the formation is generally considered to be
detrimental because it decreases the fluid efficiency (or decreases the fracture

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 14 of 27

volume created by a given volume of fracturing fluid). Excessive fluid loss can cause
early termination of a treatment due to a proppant screenout.
Fluid loss to the formation is a filtration process which is controlled by the following
parameters.
• fracturing fluid composition
• flow rate and pressure
• reservoir properties (such as permeability, porosity, pressure and fluid saturation)
• presence of microfissures, macrofissures, or faults.
During the fracturing treatment, fluid leaks-off and enters the pore spaces of the
rock. Pore-size distribution for the rock matrix varies from formation to formation.
Generally, the lower permeability formations have smaller pore openings. A 0.1-md
rock may have an average pore diameter of less than 1.0-µ, while a 500-md rock
may have an average pore diameter of 20-µ. The range of pore size may be quite
large.
Controlling fluid loss to fissures that intersect the main fracture is more difficult than
controlling fluid loss to the matrix because the openings to be blocked are larger.
Solid materials are used which can bridge the fractures and plug them, but their
effectiveness depends on the size of intersecting fissures.
Appendix E — Fluid Loss provides additional fluid-loss information.

3.5 Friction Reducers


Fluids such as water and low-viscosity oil achieve turbulence when pumped through
small tubulars at high rates. This creates high friction pressures. Dramatic
decreases in friction pressure are observed when turbulence is suppressed by
adding polymers (friction reducers) to the fluid.
Friction reducers offer no advantage unless the fluid is in turbulent flow. An already
viscosified water- or oil-base fracturing fluid will not benefit from the addition of
friction reducers. Similarly, a viscous oil pumped at low rate through casing or large
diameter tubing exhibits little opportunity for enhanced friction reduction by adding
friction reducers. High turbulence must be a factor before friction reducers are
effective.
Friction Reducers for Water
Polymers are high-molecular-weight molecules that have an affinity for water
molecules. The polymer deters turbulence by controlling migration of the individual
water molecules.
Low concentrations (10 to 20 lbm/1000 gal) of guar or HPG polymers and
copolymers of polyacrylamide are the most efficient friction reducers for water.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 15 of 27

Friction Reducers for Oil


The friction reducer for oil is an acrylonitrile copolymer. Friction reduction of as
much as 55% may be obtained using Oil Friction Reducing Agent J257 in crude oil.
Used in refined oils, friction reduction of as much as 70% may be obtained. The oil
viscosity if unaffected by friction reducers.
Friction in oil may also be reduced with low concentrations of an aluminum
phosphate ester gelling agent and an activator.

3.6 Temperature Stabilizers


Stabilizers are used to prevent degradation of water-base fracturing fluids at
temperatures greater than 200°F (93°C). The temperature stability of a fracturing
fluid is dependent on the following:
1. The stability of the polymer. Guar is less stable than HPG.
2. The fluid pH value. Guar and guar derivatives are hydrolyzed at a low pH value,
especially at elevated temperature. A high-pH fluid should be used to enhance
long-term fluid stability.
3. The presence of breakers. Fracturing fluids are degraded by breakers.
Methanol K46 and High-Temperature Gel Stabilizer J353 (sodium thiosulfate) are
stabilizers used by Dowell. Stabilizer J450 is also used in the YF100HTD, YF600HT,
and YF600UT fluids to extend the maximum temperature limit.
The reaction mechanism of methanol and sodium thiosulfate is not fully understood.
It is believed they act as oxygen scavengers and prevent rapid gel degradation
caused by dissolved oxygen. Since the dissolved oxygen content is not high, they
probably also enhance fluid stability by reacting with free radicals generated from
thermal degradation of the polymer.
The principle advantage of sodium thiosulfate is its superior performance as a
stabilizer. Normally, this performance results in the use of lower polymer
concentrations to achieve equivalent rheological properties. A comparison of
stabilizer efficiency can be made by plotting apparent viscosity versus time. Fig 1
shows the performance of a 40 lbm/1000 gal crosslinked fluid containing 50 gal
K46/1000 gal and a 40 lbm/1000 gal crosslinked fluid containing 10 lbm J353/1000
gal at 250°F (121°C) and a shear rate of 170 sec-1. J353 reduces the rate of
viscosity loss normally observed with fluids containing methanol. Fig. 2 shows the
performance of a 50 lbm/1000 gal crosslinked fluid containing 10 lbm J353/1000 gal
compared to a 60 lbm/1000 gal crosslinked fluid containing 50 gal K46/1000 gal at
275°F (135°C) and a shear rate of 170 sec-1. This comparison shows that lower
polymer loads can be used when J353 is used as a stabilizer.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 16 of 27

Fig. 1. Apparent viscosity of 40 lbm/1000 gal crosslinked fluids with methanol and
sodium thiosulfate stabilizers.

Fig. 2. Apparent viscosity of a 50 lbm/1000 gal crosslinked fluid containing sodium


thiosulfate and a 60 lbm/1000 gal crosslinked fluid containing methanol.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 17 of 27

High-Temperature Gel Stabilizer J353


J353 is more effective than K46, increasing the fracturing fluid viscosity at elevated
temperature 2- to 10-fold depending on the temperature and time of exposure to
temperature. The J353 concentration is normally 10 lbm/1000 gal in water-base
fracturing fluids.
Methanol K46
The original use of methanol in water-base fluids was for the purpose of lowering
surface tension. This was (is) desirable, especially in low-pressure reservoirs, for
enhanced fracturing fluid cleanup. These laboratory studies also revealed that
methanol enhanced the viscosity of fracturing fluids at high temperature. As a result
of these studies, methanol was introduced as a viscosity stabilizer.
K46 is normally used at a concentration of 50 gal/1000 gal in fracturing fluids. At this
concentration, K46 reduces the rate of viscosity loss, but does not significantly
reduce surface tension. K46 added to the fluid at a concentration of 50 gal/1000 gal
will reduce the surface tension by 15% to approximately 60 dynes/cm. A surfactant
added to the fracturing fluid at a concentration of 2 gal/1000 gal will reduce the
surface tension by 70% to approximately 20 dynes/cm. Fracturing fluid stability
rather than surface tension reduction is the principle advantage of K46.
K46 is inflammable, toxic, and hazardous to handle. J353 is noninflammable and
nontoxic. The use of J353 also eliminates the logistics problems when large
volumes of K46 are required. A hydraulic fracturing treatment using 400,000 gal of
stabilized fluid will require 20,000 gal of K46, but only 4000 lbm of J353.

3.7 Surfactants
A surface-active agent (surfactant) is a material which, at low concentration, adsorbs
at the interface between two immiscible substances. The immiscible substances
may be two liquids, such as oil and water, or a liquid and a gas, or a liquid and a
solid. The surfactant becomes involved in the interface and lowers the amount of
energy required to expand the interface.
Surfactants behave this way because of their unique structures. They each contain
a portion that is strongly attracted to the solvent and a portion that is not attracted to
the solvent. In the case of water, the attracted portion is called hydrophilic and the
unattracted portion is called hydrophobic. The hydrophobic portion is generally a
hydrocarbon chain that is soluble in oil but is virtually insoluble in water. The
hydrophilic portion is a very polar, often charged, group that is water soluble. If the
charge of the hydrophilic group is positive, the surfactant is cationic and if the charge
is negative, the surfactant is anionic. The surfactant is nonionic if the hydrophilic
portion is not charged.
Surfactants are often represented as in Fig. 3 and orient at an interface (for example,
between oil and water).

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 18 of 27

Fig. 3. Surfactant orientation.


Surfactants are used in fracturing fluids for various reasons. These are:
• Surfactants promote the formation of stable bubbles in foams.
• Surfactants stabilize the water-external emulsion in polyemulsion fluids. The
surfactant also acts to break the emulsion by adsorbing on the rock.
• Surfactants promote fracturing fluid cleanup by conditioning the formation and
reducing interfacial/surface tension.
Interfacial tension is a force with a dimension of dynes/cm that is a measure of the
work required to increase the surface area between two immiscible liquids by one
square centimeter. Surface tension is the same as interfacial tension except it
usually applies to gas-liquid interfaces.
Surfactants are also used to treat various types of damage. These are:
• Surfactants remove blockage by fines. Fines can be clays, silts (clay-type
minerals), or drilling-fluids solids. If a surfactant is in the fracturing fluid and wets
the individual fines particles, the particles can be removed from the formation
more easily when the fracturing fluid is cleaned-up.
• Surfactants prevent or treat near-wellbore water blocks. Surfactants reduce the
capillary pressure by lowering the surface tension of the water. Lower energy is
required to move the water through the formation matrix.
Some bactericides, clay stabilizers, and nonemulsifiers are surfactants.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 19 of 27

Wettability
The ionic charge of surfactants is important in terms of wettability. Wettability is a
term used to indicate whether the rock is preferentially coated with oil and water or
both. Almost all reservoirs are naturally water-wet, which favors oil movement
through the rock.
The contact angle is the angle between the rock surface and the fluid measured
through the fluid phase. Contact angles less than 90° measured through the water
phase indicate preferentially water-wet conditions. Contact angles greater than 90°
measured through the water phase indicate preferentially oil-wet conditions. A 90°
contact angle indicates neutral wettability; the rock surface has equal preference for
water and oil. Contact angles near 90° exhibit moderate wetting preference and
cover a range termed intermediate wettability. When the contact angle is greater
than 90°, the rock is considered non-wetting.
Cleanup is improved by lowering the capillary pressure. The capillary pressure is the
difference in pressure between a continuous oil phase and a continuous water phase
in a reservoir rock. The magnitude of this pressure difference depends on the
interfacial tension, pore space geometry, rock wettability, and the quantity of each
phase present. can be used to explain the consequence of contact angle on
capillary pressure. If a contact angle of 90° could be attained, the capillary pressure
could be reduced to near zero.
2σ cos θ
ρ= (1)
rp

Where:
ρ = capillary pressure
σ = surface tension
θ = contact angle
rp = pore radius.
Fig. 4 illustrates the wettability of oil/water/rock.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 20 of 27

Fig. 4. The wettability of oil/water/rock.


Surfactants can be adsorbed on the rock, altering the wettability. Surfactants can
also replace previously adsorbed surfactants and give the rock the wetting
characteristic of the stronger surfactant. Mixing cationic surfactants and anionic
surfactants is not advisable because of the possibility of forming precipitates. Table
1 contains the properties of common Dowell surfactants.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 21 of 27

Table 1. Properties Of Common Dowell Surfactants


Agent Nature Solubility in Surface Wettability Interfacial Tension++
Tension+ (dynes/cm)
(dynes/cm)
Brine Acid Oil Sandstone Limestone
F40 N S S S WW WW 31.6 2.5
F52.1 A S S I WW WW 24.0 <0.1
F75N N S S D WW WW 21.7 <;0.1
F78 C S S D WW WW 28.2 0.7
M38B C-N S D D OW WW 30.2 0.2
M76 C D D D WW WW 30.3 4.1
Legend:
A - Anionic
C - Cationic
D - Dispersible
I - Insoluble
N - Nonionic
OW - Preferentially oil-wet
S - Soluble
WW - Preferentially water-wet
+Surfactant concentration 1.5 gal/1000 gal in 2% KCl solution
++Aliphatic hydrocarbon — 2% solution fluid interface, surfactant concentration 1 gal/1000 gal

Table 2 illustrates the effects of wettability change. Changing from water-wet to


neutral-wet increases the relative permeability to water, but decreases the relative
permeability to oil.

Table 2. The Effects of Wettability Change


Effective Oil Permeability at Initial Contact Angle (degrees)
Water Saturation (md)
561 0
472 47
459 90
380 138
357 180

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 22 of 27

3.7.1 Fluorocarbon Surfactants


Fluorocarbon surfactants are very similar to hydrocarbon surfactants, except that the
hydrogen atoms in the hydrophobe (oil-soluble) portion are replaced by fluorine
atoms. The hydrophilic portion is essentially unchanged. Fluorocarbon surfactants
are typically much more surface-active than hydrocarbon surfactants and therefore
tend to yield lower surface tension at equal concentrations. In addition to reducing
the surface tension, fluorocarbon surfactants adsorb onto the wall of the pore space
and cause the contact angle to approach 90°. The wetting phase fluid, usually
water, can be produced from the formation more easily by reducing the capillary
pressure to near zero.

3.7.2 Surfactant Selection


Surfactant selection should be based on laboratory testing or field experience (well
performance). Laboratory testing includes compatibility testing and testing for the
critical micelle concentration. The critical micelle concentration is a narrow
concentration range in which surfactant molecules begin to aggregate and form
micelles. Adding additional surfactant will not appreciably change the interfacial
properties of the solution. The critical micelle concentration determines the amount
of surfactant required to obtain maximum reduction in surface tension.
Table 3 provides a summary of surfactant action on mineral surfaces.

Table 3. Summary of Surfactant Action on Mineral Surfaces


Mineral pH Anionic Surfactants Cationic Surfactants
Quartz <1 adsorb, oil wet do not adsorb, leave water
wet
Feldspars >2 do not adsorb, leave water wet adsorb, oil wet
<1 adsorb, oil wet do not adsorb, leave water
wet
Kaolinite 1-8
>8 do not adsorb, leave water-wet adsorb, oil wet
Illite >0 do not adsorb, leave water-wet adsorb, oil wet
Bentonite
Carbonites <8 adsorb, oil wet do not adsorb, leave
water-wet
>10 do not adsorb, leave water-wet adsorb, oil wet

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 23 of 27

3.8 Nonemulsifying Agents


An emulsion is a dispersion of two immiscible phases such as oil-in-water or water-
in-oil. The emulsion may be stabilized by formation fines or asphaltenes. Many
types of emulsifying agents occur naturally in crude oils. When formation fluids are
contacted by stimulation fluids, emulsions of varying degrees of stability may result.
The prevention or breaking (or both) of these emulsions may be accomplished by
adding nonemulsifying agents to the stimulation fluid. These agents are designed to
prevent or break emulsions formed in a reservoir.
During the stimulation treatment of an oil well, an emulsion of the fluid in the crude oil
can be formed as the stimulation fluid is being forced into the formation. This
viscous emulsion is slow to return to the wellbore and often is never returned,
especially in low-pressure wells. When this happens, the emulsion remains in the
reservoir, permanently blocking flow channels. Occasionally, the spent acid resulting
from an acid treatment is returned to the well in the form of an emulsion in crude oil.
Water-in-oil emulsions will not disperse in water but are readily dispersed by
kerosene. Similarly, oil-in-water emulsions will not disperse in oil, but will disperse in
water. Mixed emulsions or double emulsions are rarely encountered, are very
difficult to identify, and are difficult to break.

3.8.1 Nonemulsifying Agent Selection


Table 4 provides the properties of common Dowell nonemulsifying agents. Emulsion
and Sludge Preventer W35 and Nonemulsifying Agent W54 are the most commonly
used nonemulsifying agents. Silicate Control Additive M38B is sometimes used as a
nonemulsifying agent.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 24 of 27

Table 4. Properties Of Common Dowell Nonemulsifying Agents


Agent Ionic Solubility in Surface Wettability Interfacial Tension+
Nature Tension+ (dynes/cm
(dynes/cm) )
Brine Acid Oil Sandstone Limestone
W27 C S S D WW WW 29.5 4.3
W35 A-N S D D WW OW 33.8 4.9
W39 C S D D WW WW 38.2 0.7
W53 N D D D WW WW 24.9 <0.1
W54 N D D D WW WW 23.5 <0.1
Legend:
A - Anionic
C - Cationic
D - Dispersible
I - Insoluble
N - Nonionic
OW - Preferentially oil-wet
S - Soluble
WW - Preferentially water-wet
+Surfactant concentration 1 gal/1000 gal

Nonemulsifying Agent Compatibility Testing


Laboratory testing is essential in order to prevent or break an emulsion. Several
concentrations of prospective nonemulsifying agents should be tested to determine
the optimum concentration. Excessive concentrations of nonemulsifying agents can
be detrimental. The following procedure may be used for testing the compatibility of
nonemulsifying agents. This procedure is not a substitute for thorough laboratory
testing.
1. Put 50 mL of stimulation fluid (with all additives) containing a nonemulsifying
agent in a graduated container.
2. Add 50 mL of crude oil (with formation fines, if possible) to the graduated
container containing the stimulation fluid.
3. Shake vigorously for 15 sec.
4. Record the time for the fluid to break out.

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 25 of 27

4 Additive Selection
Table 5 is a recommendation guide for fracturing fluid additives. Table 6 is a
selection guide for additives. These guides are general in nature and should always
be used with caution. Fracturing fluids and additives must be engineered with the
particular reservoir in mind and with consideration for the desired performance of the
fracturing treatment.
Table 5. Additive Recommendation Guide
Bactericide Breaker Clay Fluid- Temperature Surfactan
Stabilizer Loss Stabilizer t
Additive
WF100/200 AR AR OR NLR NR AR
YF100/200 AR NLR OR NLR OR AR
YF100HTD AR NLR OR NLR AR>250°F AR

YF300LPH AR NLR OR NLR AR>200°F AR

YF400LPH AR NLR OR NLR AR>200°F AR

YF500HT AR OR OR NLR AR AR
YF600LT AR NLR OR NLR AR AR
YF600HT AR OR OR NLR AR AR
YF600UT AR NR OR NLR AR AR
StableFOA AR OR NR NLR NR AR
M (Foamer)
SuperFOA AR NLR OR NLR OR AR
M
Super AR NLR OR NLR OR AR
Sandfrac K-
1
YF"GO" III NR NLR NR NLR NR NR
YF"GO"IV NR OR NR NLR NR NR
Legend:
AR - Always recommended
NLR - Normally recommended
NR - Not recommended
OR - Occasionally recommended

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix C - Additives Dowell
Page 26 of 27

Table 6. Additive Selection Guide


Additive Type Dowell Code Compatible Fluids Recommended
Concentration
Bactericides Dryocide All water-base fluids 0.5 lbm/1000 gal
M76 All water-base fluids 0.5 to 1.5 gal/1000 gal
M275 All water-base fluids 0.3 to 0.6 lbm/1000 gal
Breakers J134L YF300/400LPH (125°F 0.25 to 1.0 lbm/1000 gal
[52°C])
WF fluids 0.0 to 1.0 lbm/1000 gal
J218 All water-base fluids 0.0 to 10 lbm/1000 gal
J297 WF fluids (<100°F 1 to 5 lbm/1000 gal
[<38°C])
J475 All water-base fluids 0 to 20 lbm/1000 gal
J479 All water-base fluids 0 to 20 lbm/1000 gal
J481 YF100HTD only 0 to 10 lbm/1000 gal
J603 YF"GO"III 0 to 5 lbm/1000 gal
M3 YF"GO"IV 0 to 50 lbm/1000 gal
Breaker Aid J318 YF100/200 (<125°F 1.0 gal/1000 gal
[<52°C])
J466 YF100D/200D (<125°F 0.5 to 2.0 lbm/1000 gal
[<52°C])
Clay Stabilizer L237 All water-base fluids 3 to 7 gal/1000 gal
L55 All water-base fluids 1 to 6 gal/1000 gal
Fluid-Loss Additives J84 All fluids 10 to 35 lbm/1000 gal
J110 All water-base fluids 25 to 50 lbm/1000 gal
J126 Ungelled oil 25 to 50 lbm/1000 gal
J168 All water-base fluids 20 to 100 lbm/1000 gal
J237 All water-base fluids 20 to 30 gal/1000 gal
J238 All water-base fluids 20 to 50 lbm/1000 gal
J418 All fluids 10 to 25 lbm/1000 gal
J426 All water-base fluids 25 to 300 lbm/1000 gal
J451 All water-base fluids 5 gal/1000 gal
J478 All water-base fluids 25 to 50 lbm/1000 gal
S100 All fluids 0.25 to 1.0 lbm/gal
Friction Reducers J166 Fresh water or brine 1 to 10 lbm/1000 gal
J257 Crude oil 10 gal/1000 gal
Refined oil
J313 Fresh water 0.25 to 1 gal/1000 gal
KCl water

Nonemulsifying W27, W39, All water-base fluids 1 to 5 gal/1000 gal


Agents
W53, W54,
W35 YF300/400 LPH 1 to 5 gal/1000 gal
WF fluids
Surfactants F75N All water-base fluids 1 to 4 gal/1000 gal
TFA 380 All water-base fluids 3 to 5 gal/1000 gal

DOWELL CONFIDENTIAL
Section 1300
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix C - Additives
Page 27 of 27

Table 6. Additive Selection Guide (continued)


Foaming Agents F40 All fluids 5 to 10 gal/1000 gal
F52.1 StableFOAM 5 to 8 gal/1000 gal (aqueous
phase)
SuperFOAM 5 to 8 gal/1000 gal (aqueous
phase)
YF300LPH 6 gal/1000 gal
YF400LPH 6 gal/1000 gal
Temperature J353 YF300/400LPH (>200°F 20 lb/1000 gal
Stabilizers [>93°C])
YF500HT/600LT or HT 10 lb/1000 gal
J450 YF100HTD 1 to 3 gal/1000 gal
YF600HT 1 to 1.3 gal/1000 gal
K46 YF100D/200D (>200°F 50 gal/1000 gal
[>93°C])
YF400LPH (>200°F 200 gal/1000 gal
[>93°C])

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 1 of 16

APPENDIX D - PROPPANTS

1 Introductory Summary............................................................................................................. 2

2 Physical Properties of Proppants .......................................................................................... 3


2.1 Proppant Strength................................................................................................................ 3
2.2 Grain Size and Grain-Size Distribution ................................................................................ 4
2.3 Quantities of Fines and Impurities ....................................................................................... 5
2.4 Roundness and Sphericity ................................................................................................... 6
2.5 Proppant Density ................................................................................................................. 6

3 Classes of Proppants .............................................................................................................. 6


3.1 Sand..................................................................................................................................... 6
3.2 Resin-Coated Proppants...................................................................................................... 6
3.2.1 Precured Resin-Coated Proppants............................................................................. 7
3.2.2 Curable Resin-Coated Proppants............................................................................... 7
3.2.3 Limitations Associated With Resin-Coated Proppants................................................ 7
3.2.3.1 Oxidative Breakers........................................................................................... 8
3.2.3.2 Borate-Crosslinked Fluids................................................................................ 9
3.2.3.3 Organometallic-Crosslinked Fluids .................................................................. 9
3.3 Intermediate-Strength Proppants......................................................................................... 9
3.4 High-Strength Proppants ..................................................................................................... 9

4 Conductivity ........................................................................................................................... 10
4.1 Closure Stress ................................................................................................................... 10
4.2 Embedment........................................................................................................................ 10
4.3 Fracture Width ................................................................................................................... 11
4.4 Proppant-Pack Porosity ..................................................................................................... 12
4.5 Factors Operative in the Presence of Polymeric Fracturing Fluids .................................... 13

5 Proppant Testing ................................................................................................................... 13

6 Measurement of Proppant Addition to Fracturing Fluids .................................................. 14

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 2 of 16

7 Proppant Selection.................................................................................................................16

8 Proppant Flowback ................................................................................................................16

FIGURES

Fig. 1. Strength comparisons of various types of proppants. .......................................................4


Fig. 2. The effect of feldspar contamination on conductivity.........................................................5
Fig. 3. The relationship of proppant concentration and fracture width (no embedment). ..........12

TABLES

Table 1. Proppant-Pack Porosity Of Sand And Intermediate-Strength Proppant .......................12


Table 2. Density Table For Proppant Added To Fracturing Fluid ...............................................14

1 Introductory Summary
Proppants are used to keep the walls of the fracture apart so that a conductive path
to the wellbore is retained after pumping has stopped and the fracturing fluid has
leaked-off. Placing the appropriate concentration and type of proppant in the
fracture is critical to the success of a hydraulic fracturing treatment.
Factors affecting the fracture conductivity (a measurement of how well a propped
fracture is able to convey the produced fluids over the producing life of the well) are:
• proppant type
• physical properties of the proppant
• proppant concentration
• proppant-pack permeability
• effects of postclosure polymer concentration in the fracture
• movement of formation fines in the fracture
• long-term degradation of the proppant.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 3 of 16

2 Physical Properties of Proppants


The physical properties of proppants that have an impact on fracture conductivity
are:
• proppant strength
• grain size and grain-size distribution
• quantities of fines and impurities
• roundness and sphericity
• proppant density.

2.1 Proppant Strength


To open and propagate a hydraulic fracture, the in-situ stresses must be overcome.
After the well is put on production, the same stresses work to close the fracture and
act on the proppant. If the proppant strength is inadequate, the closure stress will
crush the proppant and the resulting fines will plug the proppant pack. Permeability
and, therefore, conductivity of the proppant pack will be drastically reduced.
Proppants are available in different types and mesh ranges to meet the conductivity
requirements of the fracture design.
Common practice is to use the difference between the bottomhole fracturing
pressure and bottomhole producing pressure to calculate the maximum effective
stress (or closure stress) on the proppant. During flowback and testing operations,
the bottomhole producing pressure is usually held constant and at a low value to
maximize the production rate. The potential for maximum crushing can occur during
flowback and testing operations when the flowing pressure at the perforations may
be low, or initially in the production of a well because the fracture gradient is at its
maximum, decreasing with reservoir pressure depletion. However, if the well is
initially completed and produced at a higher bottomhole pressure and with a near
constant production rate, the maximum effective stress on the proppant is less. By
producing a well in this manner, the stress on the proppant can increase with time,
but never exceeds the bottomhole fracturing pressure. A higher strength proppant
can be used as a "tail-in" segment after the fracture has been packed with a lower
strength proppant as a preventive measure against induced high closure stresses
and stress concentrations near the wellbore.
Strength comparisons are shown in Fig. 1. The following general guidelines may be
used to select proppants.
• Sand — closure stresses less than 6000 psi.
• Resin-Coated Proppants — closure stresses less than 8000 psi.
• Intermediate-Strength Proppants — closure stresses greater than 5000 psi but
less than 10,000 psi.
• High-Strength Proppants — closure stresses greater than 10,000 psi.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 4 of 16

Proppant type and size should be determined by comparing economic benefits


versus cost. This is accomplished using the FracNPV* software.

Fig. 1. Strength comparisons of various types of proppants.

2.2 Grain Size and Grain-Size Distribution


Grain Size — Proppants with larger grain sizes provide a more permeable pack;
however, their use must be evaluated in relation to the formation that is propped and
the increased difficulties encountered in proppant transport and placement. Dirty
formations, or those subject to significant fines migration, are poor candidates for
large proppants. The fines tend to invade the proppant pack, causing partial
plugging and a rapid reduction in permeability. In these cases, smaller proppants,
which resist the invasion of fines, are more suitable. Although they offer less initial
conductivity, the average conductivity over the life of the well will be higher and will
more than offset the initial high productivity provided by larger proppants (which is
often followed by a rapid production decline).
Larger grain sizes can be more difficult to use in deeper wells because of greater
susceptibility to crushing due to higher closure stresses (as grain size increases,
strength decreases) and placement problems. Placement problems are two-fold —
a wide fracture is required for the larger grains, and the particle settling rate
increases with increasing size.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 5 of 16

Grain-Size Distribution — If the grain-size distribution is such that the mesh range
contains a high percentage of the smaller grains, the proppant-pack permeability and
therefore conductivity will be reduced.
Minimizing the mesh range will increase the permeability. For naturally occurring
sand, this will result in a large amount of waste. A manufactured proppant such as
intermediate-strength proppant (ISP) can be manufactured in narrow mesh ranges.
Typically, 20/40-mesh ISP is in fact nearer to 20/30 mesh.

2.3 Quantities of Fines and Impurities


Grain-size distribution and the quantities of fines and impurities in the proppant are
closely related. A high percentage of fines or impurities can reduce the proppant-
pack permeability. The effect on the proppant pack is the same as invading
formation fines. Fig. 2 illustrates the effect of feldspar contamination on conductivity.
Acid solubility is generally used as an indication of the amount of impurities such as
carbonates, feldspar, and iron oxides present in the proppant. For proppant mesh
sizes 6/20 through 30/50, the maximum allowable solubility is 2% (American
Petroleum Institute RP 56).

Fig. 2. The effect of feldspar contamination on conductivity.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 6 of 16

2.4 Roundness and Sphericity


The roundness and sphericity of a proppant grain can have a dramatic effect on
fracture conductivity. Proppant grain roundness is a measure of the relative
sharpness of grain corners, or of grain curvature. Particle sphericity is a measure of
how close the proppant particle or grain approaches the shape of a sphere. When
the grains are round and about the same size, stresses on the proppant are more
evenly distributed, resulting in higher loads before grain failure occurs.

2.5 Proppant Density


Proppant density has an influence on proppant transport and placement. High-
density proppants are more difficult to suspend in the fracturing fluid and to transport
in the fracture. Placement can be achieved in two ways — using high-viscosity fluids
which carry the proppant for the entire length of the fracture with minimal settling,
and using low-viscosity fluids at a higher flow rate. Clearly, higher-density proppants
also require more mass of material to create the same fracture volume.

3 Classes of Proppants

3.1 Sand
Sand is the most often used proppant. It is the most economical, is readily available,
and provides sufficient fracture conductivity at closure stresses less than 6000 psi.
Depending on the overall balance of physical properties, sand can be subdivided into
groups.
• Northern White Sand
• Texas Brown Sand
• Colorado Silica Sand
• Arizona Silica Sand
Based on the American Petroleum Institute (API) standards, any sand source can be
qualified and grouped similar to the above sands.

3.2 Resin-Coated Proppants


Resin coatings are applied to sand (usually Northern White Sand) to improve
proppant strength. Resin-coated sand is stronger than conventional sand and may
be used at closure stresses less than 8000 psi, depending on the type of resin-
coated sand. At closure stresses greater than 4000 psi, resin-coated sand has a
higher conductivity than conventional sand. The resin helps spread the stress over a
larger area of the sand grain and reduces the point loading. When grains crush, the
resin coating helps encapsulate the crushed portions of the grains and prevents
them from migrating and plugging the flow channel. In some cases, resin-coated
proppants may be used as an alternative to an intermediate-strength proppants.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 7 of 16

3.2.1 Precured Resin-Coated Proppants


The resin coating on the proppant is (at least partially) cured during the
manufacturing process to form a nonmelting inert film. Proppants processed in this
fashion are called precured resin-coated proppants. The major application of
precured resin-coated proppants is to enhance the performance of sand at high
stress levels.

3.2.2 Curable Resin-Coated Proppants


A curable resin coating may also be applied to sand. The major application of
curable resin-coated proppants is as an attempt to prevent the flowback of proppants
near the wellbore. Theoretically, the curable resin-coated proppants are mixed and
pumped in the last stage of the treatment and the well is shut in for a period of time
to allow the resin to bind proppant particles together and cure into a consolidated,
but permeable filter. Treatment Design provides additional information concerning
proppant flowback control methodology. Curable resin-coated proppants should only
be used as a last resort to control proppant flowback.

3.2.3 Limitations Associated With Resin-Coated Proppants


Various resin-coated proppants can affect fracturing fluid performance, and
fracturing fluids can affect the performance of resin-coated proppants. The following
guidelines are recommended if resin-coated proppants are required by the client.
• Minimize the amount of resin-coated proppants.
• Use precured resin-coated proppants. These materials are (at least partially)
cured and are typically less damaging than fully curable resin-coated proppants.
• Avoid using resin-coated proppants in conjunction with oxidative breakers.
Resin-coated proppants will decrease breaker effectiveness. Some resin-coated
proppants have very little effect on breaker activity. Additional breaker must be
used when other resin-coated proppants are used.
• Always determine fracturing fluid/resin-coated proppant compatibility before
pumping. Typically, 30 to 60% of the resin coating is lost to the fluid. Resin loss
does not behave in a straight-forward manner with changes in temperature or
time at temperature. Resin loss increases as exposure to shear increases.
• Never batch-mix resin-coated proppants in fracturing fluids.
• Minimize handling of the resin-coated proppants to keep dust levels low. Solid
resin is a good emulsion stabilizer and can create an emulsion with the fracturing
fluid that is not miscible in water.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 8 of 16

3.2.3.1 Oxidative Breakers


Breaker J218, EB-Clean ∗ J475 Breaker, EB-Clean LT J479 Encapsulated Breaker,
and Breaker J481 react with the resin coating on curable and precured resin-coated
proppants. The reaction results are:
• additional resin loss
• compressive strength reduction
• conductivity reduction
• consumption of breaker by the resin.
The particulate form of the encapsulated breakers (J475 and J479) and the resultant
localized release create a locally very high concentration of breaker. This effect has
not been extensively studied; however, numerous conductivity tests have been
performed using various fluid/breaker/resin-coated proppant combinations.
Observation of proppant-packs following fracture conductivity testing has shown that
the effect is limited to the faces of the proppant grains immediately near the site
where active breaker is being released from the breaker particle. Approximately 4 to
8 of the nearby proppant grains are affected, with roughly the closest 1/3 to 1/4
diameter (20/40 mesh) of resin coating removed from these proppant grains.
Typically, more than half of the removed resin is dissolved with the remainder
remaining as flakes. Since this is a surface reaction, the curable resin-coated
proppants are more reactive than are the precured resin-coated proppants.
J218 also reacts with resin-coated proppants; however, since the J218 is dissolved,
the effects on the resin-coated proppants are not localized. Some of the J218 is
consumed by the resin-coated proppants. Loss in curable resin-coated proppant
compressive strength should be expected. Typically, much lower (active)
concentrations of J218 (for example, 0.5 lbm/1000 gal versus 5 lbm/1000 gal) are
used so the effects are further limited.
Some resin-coated proppants have very little affect on J481. Other resin-coated
proppants reduce J481 performance considerably and additional J481 must be used.
Two important points are:
1. The effects of resin-coating degradation (flakes and loss of strength) on
proppant-pack conductivity are negligible in virtually all field conditions. The
effects of the concentrated polymer causing a reduction in conductivity are
presumed to be greater than the effects of resin-coating degradation.
2. The major problem that must be overcome is the consumption of breaker
(encapsulated or dissolved) by the (reaction with) resin-coated proppants.
Adding additional breaker will improve conductivity, but will also degrade more of
the resin coating. At some point, diminishing returns should be expected.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 9 of 16

3.2.3.2 Borate-Crosslinked Fluids


The addition of resin-coated proppants results in a decrease in the fluid pH value.
This can result in fluid decomposition. The amount of pH value decrease is
dependent on the brand and amount of resin-coated proppants.
The resin-coated proppant/fracturing fluid compatibility should be tested prior to
pumping.
Different proppants (and even different lots of proppant from a supplier) and
proppant concentration can affect the final pH value and fluid stability.
Additional resin-coated proppant information is provided in the appropriate manual
sections in the Fracturing Materials Manual — Fluids.

3.2.3.3 Organometallic-Crosslinked Fluids


The addition of resin-coated proppants, especially curable resin-coated proppants, to
organometallic-crosslinked fracturing fluids can interfere with the crosslinking
mechanism. The major factor affecting fracturing fluid performance is the reduction
in the active metal crosslinker. The metal is complexed by the resin surface and is
not available for crosslinking the polymer. Reduction in the crosslinker varies from
15 to 40%. The resin coating also lowers the pH value.
The resin-coated proppant/fracturing fluid compatibility should be tested prior to
pumping.
Different proppants (and even different lots of proppant from a supplier) and
proppant concentration can affect the final pH value and fluid stability.
Additional resin-coated proppant information is provided in the appropriate manual
sections in the Fracturing Materials Manual — Fluids.

3.3 Intermediate-Strength Proppants


ISP is fused ceramic (low-density) proppant or sintered bauxite (medium-density)
proppant. The sintered bauxite ISP is processed from bauxite ore containing large
amounts of mullite. This is in contrast to a high-strength proppant which is
processed from bauxite ore high in corundum. ISP is used at closure stresses
greater than 5000 psi, but less than 10,000 psi.

3.4 High-Strength Proppants


High-strength proppants are sintered bauxite containing large amounts of corundum,
and is used at closure stresses greater than 10,000 psi. High-strength proppant is
the most costly proppant.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 10 of 16

4 Conductivity
Fracture conductivity is a measurement of how well the propped fracture is able to
convey the produced fluid or fluids. The physical properties of proppants will affect
fracture conductivity. Other factors that have an impact on fracture conductivity are:
• closure stress
• embedment
• fracture width
• proppant-pack porosity
• factors operative in the presence of polymeric fracturing fluids.

4.1 Closure Stress


Apart from linear variations in stresses with depth due to the gradients, in-situ
stresses also depend on the lithology of the formation, tectonic components, and
pore pressure. For calculations regarding proppants, closure stress may be
represented by Eq. 1.
CS = (PfgD) – BHPP (1)
Where:
CS = closure stress (psi)
Pfg = fracture gradient (psi/ft)
D = depth (ft)
BHPP = bottomhole producing pressure (psi).

4.2 Embedment
Embedment is a process by which the fracture closes onto the proppant pack,
reducing fracture width and conductivity. It occurs when the effective closure stress
increases as a result of the reservoir pressure depletion.
When fracturing formations where embedment can occur, a wider fracture and a
higher concentration of proppant (weight per unit of fracture area) are required to
counteract the reduction in conductivity due to embedment.
The depth of embedment will be limited to one-half of a proppant grain diameter on
each fracture face in most formations. However, in very soft formations such as
chalk, the depth of embedment may be more severe.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 11 of 16

4.3 Fracture Width


The propped fracture width (no embedment) is related to the proppant concentration
by Eq. 2.
C p 12
w=
(1 − φ p )ρ
(2)

Where:
w = propped width (in.)
Cp = proppant concentration (lbm/ft2)
φp = proppant-pack porosity (%)
ρ = proppant absolute density (lbm/ft3).
Although the maximum conductivity can theoretically be obtained from a partially
monolayer system, actual placement of a partial monolayer is virtually impossible to
achieve. Proppant embedment can cause total loss of conductivity in a monolayer
system. Therefore, the propped fracture is usually designed to have multiple layers
of proppant. By increasing the proppant concentration, multiple layers of proppant
will result and the fracture conductivity will increase because of the increased
fracture width associated with multiple layers. Multiple-layer proppant packing
requires a fracturing fluid with enough viscosity to create a fracture width that is
sufficient for entrance of the proppant at a higher concentration.
Fig. 3 illustrates the relationship of proppant concentration and fracture width for
20/40-mesh sand. Once a multilayer packing is achieved, the fracture width
increases proportionately to the increase in proppant concentration. Consequently,
the fracture conductivity also increases. The fracture width must be sufficient to
inject the proppant at a higher concentration, and the selected proppant must have
sufficient permeability under closure stress conditions to maintain the optimum
conductivity.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 12 of 16

Fig. 3. The relationship of proppant concentration and


fracture width (no embedment).

4.4 Proppant-Pack Porosity


The fracture width is reduced as the closure stress increases. This reduces the pore
volume to proppant volume ratio within the fracture, and therefore reduces the
proppant-pack porosity.
Table 1 provides the proppant-pack porosity of sand and intermediate-strength
proppants for various closure stresses.
Table 1. Proppant-Pack Porosity Of Sand And Intermediate-Strength
Proppant
Porosity (%) Closure Stress, Sand (psi) Closure Stress, ISP (psi)
37.0 1000 4000
33.5 3000 6000
30.0 5000 8000

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 13 of 16

4.5 Factors Operative in the Presence of Polymeric Fracturing Fluids


Proppant-pack permeability is significantly impaired by the concentration of polymer
in water-base fracturing fluids. The severity of damage is strongly dependent on the
type of fluid used and the surface polymer concentration. Final postclosure polymer
concentrations can be 10 to 15 times greater than the original polymer
concentration. Additional information on proppant-pack damage is provided in .

5 Proppant Testing
The API has established recommended practices for testing proppants. The
proppant evaluation includes the following:
• proppant sampling and splitting
• determination of bulk density, apparent density, and absolute density
• sieve analysis
• determination of roundness and sphericity
• determination of acid solubility
• silt test (turbidity method)
• determination of crush resistance.
The following are the testing references.
API RP 56, Recommended Practices for Testing Sand Used in Hydraulic
Fracturing Operations , First Edition, American Petroleum Institute (1983).
API RP 60,Recommended Practices for Testing High-Strength Proppants
Used in Hydraulic Fracturing Operations, First Edition, American Petroleum
Institute (1989).
These testing references also contain the minimum and maximum allowables for the
different proppant properties. The allowables are generally considered as minimum.
For example, the roundness and sphericity of 0.6 is often exceeded and may be 0.9
for "high quality" sand.
API RP 56 and API RP 60 are available from the American Petroleum Institute,
Publications and Distribution Section, 1220 L Street, NW, Washington, DC 20005,
Telephone: 202.682.8375.

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 14 of 16

6 Measurement of Proppant Addition to Fracturing Fluids


Hydraulic fracturing fluids containing proppants are slurries. Slurries are composed
of a fluid and a solid. The slurry density, or mass per unit volume of fluid, is
expressed as lbm/gal or g/cm3 or kg/m3. Slurry density is a critical parameter used
in hydraulic fracturing operations to control the amount of proppant added to the
fracturing fluid.
Occasionally, two terms are used interchangeably, but have very different meanings.
These are PPA, which is Pounds of Proppant Added to a gallon of liquid, and
lbm/gal, which is slurry weight in Pounds Per Gallon. Fracturing treatments are
designed using stages of slurry volumes containing ever increasing proppant
concentrations in the slurry. As more proppant is added to the slurry, the quantity of
liquid required decreases due to the additional volume taken up by the proppant
(illustrated in Table 2).
Table 2. Density Table For Proppant Added To Fracturing Fluid
Specific Gravity of Fracturing Fluid = 1.00
Weight Density of Fracturing Fluid = 8.345 lbm/gal
Proppant is Sand (SG = 2.65, Density = 22.1 lbm/gal)
Pounds Pounds Weight Gallons in One Slurry
Proppant Proppant Density Liquid SG Gallon Yield
Added to in one of Slurry of Slurry Slurry
One Gallon (lbm/gal)
Gallon Slurry
0.00 0.00 8.35 1.00 1.00 1.00
1.00 0.96 8.94 1.07 0.96 1.05
2.00 1.83 9.49 1.14 0.92 1.09
3.00 2.64 9.99 1.20 0.88 1.14
4.00 3.39 10.45 1.25 0.85 1.18
5.00 4.08 10.88 1.30 0.82 1.23
6.00 4.72 11.28 1.35 0.79 1.27
7.00 5.32 11.66 1.40 0.76 1.32
8.00 5.87 12.00 1.44 0.73 1.36
9.00 6.40 12.33 1.48 0.71 1.41
10.00 6.89 12.63 1.51 0.69 1.45
The properties of a fracturing fluid are controlled with the use of additives. The
additive quantities in a fracturing fluid are always calculated based on a "clean" fluid
(fluid without proppant), and not on the slurry. As Table 2 illustrates, large errors in

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix D - Proppants
Page 15 of 16

the calculation of additive quantities would take place if calculations were based on
the slurry volume rather than the liquid phase.
The slurry density in fracturing treatments is measured by a densitometer. The
densitometer will read either pounds of proppant added to one gallon of liquid (PPA),
weight density of fluid (lbm/gal), or both.
Occasionally, the slurry density weight, specific gravity of the slurry or liquid, or
proppant volume in the slurry must be calculated. The following equations may be
used for these purposes.
Pounds of Proppant in One Gallon of Slurry (Eq. 3)

χ
(3)
χ
1+
8.345SG p

Weight Density of Slurry (lbm/gal) (Eq. 4)

χ + 8.345 SGL
(4)
χ
1+
8.345 SG p

Specific Gravity of Slurry (Eq. 5)

χ + 8.345 SGL
(5)
χ
8.345 +
SG p

Gallons of Liquid in One Gallon of Slurry (Eq. 6)


χ
1− (6)
8.345SG p + χ

Slurry Yield (Eq. 7)


χ
1+ (7)
8.345 SG p

Where:
SGL = specific gravity of the liquid phase
SGp = specific gravity of the proppant
x = pounds of proppant added to a gallon of liquid (lbm).

DOWELL CONFIDENTIAL
Section 1400
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix D - Proppants Dowell
Page 16 of 16

7 Proppant Selection
The selection of a proppant involves many compromises imposed by economical
and practical considerations. Criteria for selecting proppant type, size, and
concentration is based on a match of the fracture flow capacity to the formation
permeability to provide the highest production rates consistent with economics.
Treatment Design provides proppant selection and scheduling information.

8 Proppant Flowback
Treatment Design provides information concerning proppant flowback.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 1 of 19

Appendix E - Fluid Loss

1 Introductory Summary............................................................................................................. 2

2 Filtrate-Dependent Leakoff ..................................................................................................... 3


2.1 Wall-Building Coefficient ...................................................................................................... 5
2.2 Viscosity Control Coefficient ................................................................................................ 7
2.3 Compressibility Coefficient................................................................................................... 7
2.4 Fracturing Fluid Coefficient .................................................................................................. 8
2.5 Total Leakoff Volume ........................................................................................................... 8
2.6 Fluid-Loss Mechanisms and Permeability............................................................................ 8

3 Pressure-Dependent Leakoff .................................................................................................. 9


3.1 Geologic Discontinuities....................................................................................................... 9
3.2 Opening Natural Fissures .................................................................................................... 9
3.3 Fluid-Loss Control in Fissures ........................................................................................... 10

4 Types of Fluid-Loss Additives.............................................................................................. 11


4.1 Inert Particulates ................................................................................................................ 11
4.2 Soluble Particulates ........................................................................................................... 11
4.3 Oil-in-Water Emulsions ...................................................................................................... 12

5 Formation Considerations .................................................................................................... 12


5.1 Fluid Loss to the Rock Matrix............................................................................................. 12
5.1.1 Pore-Size Determination........................................................................................... 13
5.1.1.1 Experimentally ............................................................................................... 13
5.1.1.2 Calculation ..................................................................................................... 13
5.2 Fluid Loss to Fissures ........................................................................................................ 13

6 In-Situ Measurement of Fluid-Loss Coefficients ................................................................ 16

7 Guide to Dowell Fluid-Loss Additives ................................................................................. 17


7.1 Fluid-Loss Additive Sizing .................................................................................................. 18

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 2 of 19

FIGURES

Fig. 1. Zones of invasion. .............................................................................................................3


Fig. 2. The relationship of the fluid-loss coefficient to fluid volume and fracture length. .............4
Fig. 3. Typical fluid-loss data for a wall-building fluid. ..................................................................6
Fig. 4. Log-log plot of fracture pressure indicating possible fluid loss to fissures......................10
Fig. 5. Opening of fissures..........................................................................................................10
Fig. 6. Idealized G Plot (from the DataFRAC software)..............................................................16
Fig. 7. Bridging particle size versus pore-throat diameter. .........................................................18
Fig. 8. Bridging particle size versus approximate permeability...................................................19

TABLES

Table 1. Influence Of Natural Fissures In Low-Permeability Rock .............................................15


Table 2. Guide To Dowell Fluid-Loss Additives ..........................................................................17

1 Introductory Summary
Fluid loss can be described as leakage of the fracturing fluid out of the main fracture.
The rate of fluid leakoff (fluid loss) to the formation is one of the most critical factors
involved in determining fracture geometry and treatment design. Only the fluid that
remains in the main fracture is useful in propagating the fracture. Fluid that leaks-off
is wasted. The volume of fluid loss during a hydraulic fracturing treatment
determines the fracturing fluid efficiency (ratio of fracture volume to volume
pumped). Excessive fluid loss can cause early termination of a treatment due to a
proppant "screenout." The rate of fluid loss also influences fracture closure time and
may influence proppant distribution in the fracture.
Fluid loss to the formation is a filtration process which is controlled by the following
parameters.
• fracturing fluid composition
• flow rate and pressure
• reservoir properties (permeability, porosity, pressure and fluid saturation)
• presence of microfissures, macrofissures, or faults.
The efficiency of a fracturing fluid can be greatly improved by adding materials to
control fluid loss.
Reactive Fluids
A discussion of fluid loss associated with reactive fluids is provided in Acid
Fracturing.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 3 of 19

2 Filtrate-Dependent Leakoff
Prior to the fracturing treatment, the formation contains a number of fluids (for
example, hydrocarbons, water) with different flow properties. When the fracturing
fluid penetrates a formation, there may be three zones (Fig. 1):
• a filter cake with varying thickness (R1)
• a zone invaded by the filtrate (R2)
• a region with the reservoir fluids only (R3).

Fig. 1. Zones of invasion.


For the purposes of treatment design, the comparative effectiveness of different
fracturing fluids is expressed in terms of a fluid-loss coefficient (units of ft/min½).
The fluid-loss coefficient is a measure of the flow resistance of the fracturing fluid
leaking off into the formation during pumping operations. The relationship between
the fracturing fluid characteristics and the formation characteristics is determined by
the fluid-loss coefficient.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 4 of 19

During a fracturing treatment, only the volume of fracturing fluid that remains within
the walls of the fracture is effective. The fluid that leaks-off into the formation matrix
is lost, insofar as additional fracture extension is concerned. Fig. 2 illustrates the
relationship of the fluid-loss coefficient to fluid volume and fracture length.

Fig. 2. The relationship of the fluid-loss coefficient to fluid


volume and fracture length.
The fracturing fluid effectiveness is dependent upon the various linear flow
mechanisms controlling the amount of fluid loss taking place during the fracturing
treatment.
There are three types of linear flow mechanisms (fluid-loss coefficients). These are
• the wall-building coefficient (Cw)
• the viscosity control coefficient (Cv)
• the compressibility coefficient (Cc).
The Cw is determined experimentally. The Cv and Cc can be calculated from reservoir
data and fracturing-fluid viscosity. All three mechanisms act simultaneously during a
fracturing treatment and affect the efficiency of the fluid. The rate of fracturing fluid
loss into the formation is governed by the Fluid-Loss Coefficient (C), a combination
of all three mechanisms.
Example calculations of the leakoff coefficients and leakoff comparisons are
provided in A Practical Companion to Reservoir Stimulation.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 5 of 19

2.1 Wall-Building Coefficient


Refer to Fig. 1. In the first region (R1), polymer and fluid-loss additive particles are
deposited and form a filter cake. The permeability of the cake is very low. The rate
of fluid flow through the cake is governed by Cw. The spurt value (Sp) represents the
volume of fluid that leaks off during formation of the filter cake. The spurt loss
reduces the volume of fluid available for fracture extension.
For wall-building fluids, the filter cake continues to grow with time and the fluid-loss
rate decreases. For ideal wall-building fluids, a plot of filtrate volume versus the
square root of time produces a straight line. The slope of this line is used to
calculate the Cw and the y intercept is used to calculate the Sp. The Cw is directly
proportional to the leakoff velocity through the established filter cake. By using the y
intercept to calculate spurt, the assumption is made that the filter cake is
instantaneously established. Since a finite time is required for an effective cake to
be formed, calculated spurt only approximates the fluid volume lost during filter-cake
formation.
Eq. 1 is the mathematical expression for Cw and Eq. 2 is the mathematical
expression for Sp.
0.0164m
Cw = , ft / min 1/ 2 (1)
A
Where:
m = slope (mL/min½)
A = area of the core (cm2)
24.4b
Sp = , gal / 100 ft 2 (2)
A
Where:
b = y intercept (mL)
A = area of the core (cm2)
Fig. 3 illustrates data from a typical fluid-loss test using a wall-building fluid. Cw
values vary from as low as 0.0001 ft/min½ to a maximum of 0.03 ft/min½. The
lower values indicate a more efficient fluid.
Fracturing fluids containing guar or hydroxypropylguar (HPG) polymers with
appropriate fluid-loss additives are good wall-building fluids.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 6 of 19

Fig. 3. Typical fluid-loss data for a wall-building fluid.


THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY.
Testing Under Static and Dynamic Flow Conditions
Although most fluid-loss data are generated in static conditions, these rates may be
misleading because the filter cake is allowed to grow without being subjected to
erosion. Also, the effects of shear on fluid-loss rates cannot be determined in static
tests.
Filter-cake erosion and fluid degradation under conditions of shear and temperature
have been the subject of considerable study. The results from these studies have
shown that dynamic filtration velocities of fracturing fluids tend to increase as the
shear rate and temperature increase. Dynamic flow fluid-loss tests that were
performed at 40 sec-1 (which approximates fracture conditions in many instances)
produced data similar to static test results. Comparable results were found for
uncrosslinked fluid and for borate- and organometallic-crosslinked fluids. Also, shear
rates greater than 80 sec-1 in dynamic tests caused organometallic-crosslinked fluid
to have higher leakoff rates than uncrosslinked- or borate-crosslinked fluid.
Differences in fluid-loss behavior among the fluids seemed directly related to
differences in filter cake. The organometallic-crosslinked fluid produced a thin,
consolidated filter cake about one-half as thick as the soft filter cake formed by the
uncrosslinked and borate-crosslinked fluids.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 7 of 19

2.2 Viscosity Control Coefficient


The second region (R2 in Fig. 1) contains the filtrate. Viscosity of the filtrate and
relative permeability to the filtrate tend to control fluid loss when the viscosity of the
fracturing fluid is greater than that of the reservoir fluid. The Cv is most likely to be in
effect when a gas reservoir is fractured with a nonwall-building, high-viscosity fluid.
Polyemulsions, water-base fluids containing hydroxyethylcellulose (HEC) polymers,
and viscous oils are fluids that have a high-viscosity filtrate. (The viscosity of the
filtrate from a wall-building fluid is much lower than the viscosity of the fracturing fluid
itself if the polymer filters out on the formation face.)
The filtrate flow rate is governed by Darcy's law. Derived from Darcy's law is the
mathematical expression (Eq. 3) for Cv.
1/ 2
− 3  k f ∆pφ 
Cν = 1.48 E 
 µf


(
, ft / min 1/ 2 ) (3)

Where:
kf = relative formation permeability to the filtrate (md)
∆p = differential pressure between the fracture and the formation pore pressure (psi)
φ = formation porosity (decimal fraction)
µf = viscosity of the filtrate at bottomhole conditions (cp).

2.3 Compressibility Coefficient


The third region (R3 in ) has the flow of the native fluids alone. The compressibility,
viscosity, and relative permeability to those reservoir fluids affect the rate of leakoff
of the fracturing fluid. This control mechanism is most effective when the reservoir
and fracturing fluids have very similar physical properties and are not very
compressible. Leakoff is slowed because the noncompressible fracturing fluid must
compress and displace the noncompressible reservoir fluid in order to penetrate the
formation matrix. The mathematical expression (Eq. 4) for Cc is

( )
1/ 2
−3  k c φ
Cc = 119
. E ∆p r t  , ft / min 1/ 2 (4)
 µr 
Where:
∆p = differential pressure between the fracture and the formation pore pressure (psi),
kr = formation permeability to formation fluid (md),
ct = total formation compressibility (psi-1),
φ = formation porosity (decimal fraction), and
µr = viscosity of the formation fluid at reservoir conditions (cp).

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 8 of 19

2.4 Fracturing Fluid Coefficient


The rate of fracturing fluid loss into the formation is governed by C which is a
combination of the coefficients Cw, Cv, and Cc. Several methods have been proposed
for combining Cw, Cv, and Cc. The simplest method of doing this is to conclude that
the rate of fluid loss is primarily regulated by either the effectiveness of the filter cake
(it is less permeable than the rock) or by the formation and fluid characteristics. For
this purpose, Cv and Cc are combined to form the viscosity, compressibility coefficient
(Cvc) and it is then compared to Cw. The smaller of the two is used in fracture design
calculations. This is analogous to flow being limited by the smaller of two chokes in
series. The mathematical expression (Eq. 5) for Cvc is:
2Cν Cc
Cνc = , ft / min 1/ 2 (5)
( )
1/ 2
Cν + Cν2 + 4Cc2

2.5 Total Leakoff Volume


An estimate of the total leakoff volume (V) per unit area at any time can be
calculated using Eq. 6. This equation assumes an average of the formation
permeability. A more accurate value may be obtained by calculating Cv , Cc, and Sp
for individual sections of the formation height exhibiting a range of permeabilities.

( )
V = S p + 2( Cνc or Cw ) t 1/ 2 (7. 48), gal / ft 2 ( ) (6)

Where:
Sp = spurt loss (gal/ft2)
Cvc = viscosity, compressibility coefficient (ft/min½)
Cw = wall-building coefficient (ft/min½)
t = time (min).

2.6 Fluid-Loss Mechanisms and Permeability


The fluid-loss-control mechanisms apply for all permeabilities. For instance, in low-
permeability, hydraulically fractured formations, the three regions of filter-cake,
filtrate-invaded, and reservoir-fluid zones provide the pressure differentials to control
leakoff. The depth of polymer invasion is limited to near the fracture surface and the
dominant filter cake quickly forms. In high-permeability formations, a fourth region
exists where polymer and particulates invade, providing reduced formation
permeability and additional pressure loss. This is referred to as deep-bed filtration.
Additional fluid-loss information for high-permeability formations is provided in
HyPerSTIM Service.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 9 of 19

3 Pressure-Dependent Leakoff
Pressure-dependent leakoff due to fluid loss into fissures leads to non-ideal treating
pressure behavior and is thought to contribute to screenout in low-permeability
formations where limited fluid loss would otherwise be anticipated.

3.1 Geologic Discontinuities


Fissures and other forms of discontinuities are a widespread characteristic of rock
masses. Dimensionally, discontinuities extend from the scale of crystal features (for
example, transform faults) to the scale of petrographic features (for example,
microcracks). Evidence also suggests that these features exist from near the
surface to great depths.
The extent to which the behavior of a discontinuous rock mass differs from that of a
rock mass consisting of only the host rock matrix is a function of the discrepancy
between the properties of the discontinuities and those of the matrix. For example,
in a crystalline lithology, a few small aperture fractures may result in a large-scale
permeability which is many orders of magnitude greater than that of the matrix. The
same fractures in a more permeable clastic lithology may be of little consequence.
The extent to which discontinuities contribute to rock-mass behavior is also governed
by the scale of the features relative to a scale of interest (for example, the height or
length of a hydraulic fracture).

3.2 Opening Natural Fissures


Fig. 4 illustrates a log-log plot of net pressure versus time. Small positive slope
(phase 1) indicates the fracture is propagating under contained height and free
lateral extension. Zero slope (phase 2) indicates constant pressure. One
interpretation for constant pressure is increased fracture height; however, increased
fracture height is usually indicated by negative slope. Another interpretation is the
opening and inflating of fissures that are crossed by the hydraulically induced
fracture. These fissures normally have relatively higher permeability than the matrix
and the fluid can readily penetrate into the fissures and maintain a pressure nearly
equal to the pressure in the primary fracture. The fissures will open when the fluid
pressure exceeds the formation stress acting across them. When this magnitude of
pressure is reached, the fissures open and act to regulate the pressure at this critical
magnitude (phase 2). A significant portion of the fracturing fluid can be lost in this
process because of a large number of fissures that can open at this critical pressure.
Unit positive slope (phase 3) indicates flow restriction. The flow restriction can be
caused by accelerated fluid loss, leading to excessive slurry dehydration and a
screenout.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 10 of 19

Fig. 4. Log-log plot of fracture pressure indicating


possible fluid loss to fissures.
Fig. 5 also illustrates the different phases. In phase 1, the fracture extension
pressure (pe) is less than the minimum stress of the fissure (σf) and the fracture
continues to propagate. In phase 2, pe is approximately equal to σf. The net
pressure required for the fissures to open and inflate = σf - σH,min /1-2v ≅ 1.5 ∆σ where
v is Poisson's ratio for the rock and ∆σ is the difference in principal horizontal
stresses. This implies that effective fracturing not only requires a significant stress
difference for vertical barriers, but also a significant stress difference between the
principal stresses in the horizontal plane to avoid opening fissures. In phase 3, the
flow is being restricted.

Fig. 5. Opening of fissures.

3.3 Fluid-Loss Control in Fissures


Fluid-loss additives can be used to control fluid loss to fissures. A 300-mesh
material in the pad fluid will bridge fissures before they open, but will not bridge open
fissures or the fracture tip. A 100-mesh material between the pad fluid and slurry will
bridge open fissures. Tip screenouts can also be avoided by using 100-mesh
material between the pad fluid and slurry.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 11 of 19

High-quality foam fluids can retard fluid loss in fissures because of high yield stress
in the fluid. Yield stress in foam fluids is discussed in Foam Fracturing. The use of a
foam fluid may alter the pressure response of opening fissures and provide a more
gradual transition between phase 1 and phase 2 (Fig. 4 and Fig. 5).

4 Types of Fluid-Loss Additives


A wide range of liquid and solid fluid-loss additives is available. Additives are divided
into several types.

4.1 Inert Particulates


Inert fluid-loss additives are the least costly and most widely used additives, but may
cause some reduction in formation and proppant-pack permeability and conductivity.
Although permeability damage in the formation near the fracture walls is not desired,
studies indicate that it has little effect on well productivity. Proppant-pack damage
can have a significant effect on well productivity.
Silica flour (Fluid Loss Agent J84 and Fluid-Loss Additive J418) is milled silicon
dioxide. Due to its wide range of particle sizes, it is effective in reducing fluid loss to
the rock matrix and fissures. A 10-fold reduction in spurt loss for 5- to 100-md rock
has been reported when silica flour is used. The effectiveness of inert fluid-loss
additives is enhanced if the fracturing fluid contains a material to agglutinate
(combine) with the fluid-loss additive. An agglutinative agent acts as a sealer around
the inert fluid-loss additive. Guar or HPG polymers will act as agglutinative agents in
water-base fluids. Soluble polymers such as cellulose derivatives or xanthan will not
develop a filter cake and, therefore, are not agglutinative agents. Most crude oils
contain enough paraffin or asphaltenes to act as agglutinative agents.
Adomite Aqua+ (Fluid-Loss Additive J110) is a mixture of enzyme- and temperature-
degradable components and is effective at reducing fluid loss to the rock matrix only.
J110 is the most costly of the inert fluid-loss additives.
Other inert wall-building agents include FLA* J478 Slurriable/Degradable Additive,
which is composed of starches and Adomite Mark II (Fluid-Loss Additive J126) which
is intended for use in oil-base fracturing fluids other than YF* "GO" fluids.

4.2 Soluble Particulates


Oil-soluble resins (Fluid-Loss Additive J168 and Fluid-Loss Additive J426) are
effective fluid-loss additives. These materials, when they are sized properly and
have a high enough softening point, can bridge the pore spaces and fissures and
cause plugging and reduce fluid loss. An advantage that these materials have over
inert particulates is that they are oil soluble and are dissolved in the produced
hydrocarbon. Because they dissolve, there is no formation or proppant-pack

+ Trademark of Nalco Chemical Company


* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 12 of 19

damage associated with these materials. A wide range of particle sizes is available.
These additives are more costly than the inert particulates.
Fluid Loss Agent J238 is an oil-soluble resin that can be used to control leakoff in
aqueous base fracturing fluids. Its primary use is as a fluid-loss additive in matrix
treatments.

4.3 Oil-in-Water Emulsions


An effective and popular method for controlling fluid loss is to use emulsified fluids.
Petroleum-base additives (Diesel Oil U51 and Fluid-Loss Additive J451) are used to
create oil-in-water emulsions with water-base fracturing fluids. The two-phase fluids
reduce the relative permeability in the filter cake and formation, and exhibit good
fluid-loss control. The water-base fluid often contains guar or HPG as viscosifiers
which further enhance the fluid-loss properties of the emulsion.
Laboratory testing has shown that J451 at a concentration of 5 gal/1000 gal is as
effective as U51 at a concentration of 50 gal/1000 gal. Using J451 instead of U51
will be an economic benefit for the client because the diesel cost, transportation cost,
and additional pumping cost are not incurred.
Static testing and dynamic flow testing of 5% diesel as a fluid-loss additive have
demonstrated different results. In a static test, the 5% diesel reduced the fluid-loss
rate by a factor of 5. In a dynamic flow test, the reduction was only a factor of 1.5. A
possible explanation for the difference is that the dynamic filter cake containing oil is
very thin compared to one without oil, indicating oil has a detrimental effect on filter-
cake durability and thickness.

5 Formation Considerations
The rate of fluid loss and the effectiveness of the fluid-loss materials depend on
certain formation characteristics. Loss may be to the matrix or fissures or both.

5.1 Fluid Loss to the Rock Matrix


If the fluid contains particulates of the proper size, these particulates will bridge
pores at the fracture face and enhance the formation of the filter cake. Pore size
and distribution in the rock matrix vary. Lower permeability formations generally
have smaller pore openings. A 0.1-md rock may have an average pore diameter of
less than 1.0 µm, while a 500-md rock may have an average pore diameter of 20
µm. The range of pore sizes may be quite large, which makes it beneficial to have a
wide range of particle sizes so that all pore spaces can be bridged. The ground
silica materials (J84, J418) and oil-soluble resin (J168) can be used to control fluid
loss to the matrix.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 13 of 19

The SUPER SANDFRAC* K-1 fluid is a water-in-oil emulsion (67% oil and 33%
water, stabilized with an emulsifier). Because it is an emulsion, K-1 has good fluid-
loss characteristics that can be further enhanced by the addition of J84 or J418.
Water-base fluids containing J451 or U51 are also very effective controlling fluid loss
to the rock matrix.
Foam fracturing fluids are considered wall-building fluids when polymers are part of
the liquid phase. The Cw for a foam is a combination of the liquid and gas leakoff
velocities. Cw increases with permeability while Sp is zero at all permeabilities. Cw
decreases as polymer concentration increases in the liquid phase. Foam quality
(volume of the gas divided by the total volume) has little effect on Cw in low
permeability (<1-md) reservoirs. Cw decreases as foam quality increases in
reservoirs where the permeability is >10 md.

5.1.1 Pore-Size Determination


Pore-throat diameter can be determined (1) experimentally or (2) by calculation.

5.1.1.1 Experimentally
Pore-throat diameter can be determined by mercury injection into a core sample.
This method is destructive; the core cannot be used for any further testing.
Alternatively, thin-section analysis can be performed on a small core section.

5.1.1.2 Calculation
A representative pore-throat diameter can be calculated using Eq. 7 and permeability
and porosity data. This calculated pore-throat diameter is not the average pore-
throat diameter, but is the pore-throat diameter that corresponds to the entry
capillary pressure (the expression stresses that most of the flow, and therefore, most
of the contribution to permeability, is occurring in relatively large pores).
 k 1/ 2 
d 0.475
=  (7)
 φR 
Where:
d = pore-throat diameter (micron)
k = permeability (md)
φr = porosity (decimal)

5.2 Fluid Loss to Fissures


Controlling fluid loss to fissures that intersect the main fracture is more difficult than
controlling fluid loss to the matrix because the openings to be blocked are larger.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 14 of 19

Fissures can dramatically increase the permeability of the formation. Inert materials
such as silica flour which can bridge the fissure and plug it off are used, but their
effectiveness depends on the width of the intersecting fissure. Laboratory testing to
compare the relative fluid flow through simulated fissures indicates that a fluid-loss
additive is most effective at bridging these fissures when at least 60% of the material
is composed of particles that are larger than the fissure width. Silica flour smaller
than 200 mesh (J418) is very useful for microfissures (fissures less than 50 µm
[0.002 in.] in width), but particles larger than 200 mesh (J84, 100-mesh sand) are
necessary for macrofissures (fissures greater than 50 µm [0.002 in.] in width).
Particulates in different mesh sizes are also available as oil-soluble resins. J168 is
effective in microfissures, while J426 is more effective in macrofissures.
Table 1 illustrates the influence of fissures in low-permeability rock.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 15 of 19

Table 1. Influence Of Natural Fissures In Low-Permeability Rock


Average Permeability (md) as Influenced by Fissure of
Given Width (in.)
Fissure 0.0005 0.001 0.002 0.005 0.010
Interval (ft)
0.1-md Matrix
0.5 1.23 9.10 73 1126 9000
1.0 0.66 4.60 37 563 4500
2.0 0.38 2.35 19 282 2250
5.0 0.21 1.00 8 113 900
10.0 0.16 0.55 4 57 450
20.0 0.13 0.33 2 29 225
50.0 0.11 0.19 1 12 90
100.0 0.11 0.15 1 6 45
0.5-md Matrix
0.5 1.62 9.50 73 1126 9000
1.0 1.06 5.00 37 563 4500
2.0 0.78 2.75 19 282 ,250
5.0 0.61 1.40 8 113 900
10.0 0.56 0.95 4 57 450
20.0 0.53 0.73 2 29 225
50.0 0.51 0.59 1 12 90
100.0 0.51 0.55 1 6 45
1.0-md Matrix
0.5 2.13 10.00 73 1126 ,001
1.0 1.56 5.50 37 564 4501
2.0 1.28 3.25 19 282 ,251
5.0 1.11 1.90 8 114 901
10.0 1.06 1.45 5 57 451
20.0 1.03 1.23 3 29 226
50.0 1.01 1.09 2 12 91
100.0 1.51 1.05 1 7 46

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 16 of 19

The influence of fissures on the average permeability of the fracture face is


determined using Eq. 8 and demonstrated in the example.

( 4.54)(10 9 )( w 3 )
k avg = k md + (8)
L
Where:
kavg = average permeability of fracture face (md)
kmd = primary permeability of matrix (md)
w = microfracture width (in.)
L = distance between microfractures (ft).
Example
A core test indicates that a particular formation has a permeability of 0.1 md, but
pressure-buildup tests indicate that the formation has an average permeability
between 0.3 to 0.4 md. The formation is obviously producing from fissures. Data in
indicate the possible combinations of microfissure width and interval that would
produce this much difference in measured permeability. This formation could
possibly contain one 0.0005-in. fissure every 2.0 ft along the fracture face, or a
0.001-in. fissure every 20 ft.

6 In-Situ Measurement of Fluid-Loss Coefficients


As part of the calibration treatment, the bottomhole pressure is monitored and
recorded until fracture closure is attained. The data is analyzed using the G Plot in
the DataFRAC* module of the FracCADE* software (Fig. 6). The slope of the G Plot
and compliance determine the fluid-loss coefficient. DataFRAC Service provides
additional discussion on pressure decline analysis following a calibration treatment.

Fig. 6. Idealized G Plot (from the DataFRAC software).

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 17 of 19

7 Guide to Dowell Fluid-Loss Additives


Table 2 provides a guide to Dowell fluid-loss additives. This selection guide should
be considered as general recommendation. Fluid-loss additives must be engineered
considering the reservoir characteristics and the desired performance of the
treatment.
Table 2. Guide To Dowell Fluid-Loss Additives
Additive Composition Solubility Carrier Fluid Recommended Use
and Quantity
J66S Fine Salt Water Oil-base fluids or Designed for
saturated brine macrofissures. 0.25 to
2 lbm/gal
J84 Silica Flour Inert Water-base or oil-base Matrix, microfissure and
fluids macrofissure control, 10
to 35 lbm/1000 gal
J110 Colloids, Inert Water-base or oil-base Matrix control, 25 to 50
Polymers fluids lbm/1000 gal
J126 Aluminosilicate Inert Oil-base fluids Matrix and microfissure
and Fatty Acid (uncrosslinked) control, 25 to 50
lbm/1000 gal
J168 Hydrocarbon Oil Water-base fluids Matrix and microfissure
Resins control, 20 to 100
lbm/1000 gal
J237A Liquid, Oil Water-base fluids 30 to 50 gal/1000 gal
Nonionic
Resins
J238 Hydrocarbon Oil Water-base fluids Matrix control, 20 to 50
Resins lbm/1000 gal
J330 Liquid-Anionic Oil Water-base fluids 6 to 24 gal/1000 gal
Resins
J418 Silica Flour Inert Water-base or oil-base Matrix and microfissure
fluids control, 10 to 25
lbm/1000 gal
J426 Hydrocarbon Water-base Oil Macrofissure control, 25
Resins fluids to 300 lbm/1000 gal
J451 Proprietary Solvent Water-base guar or HPG Matrix control, 5
crosslinked fluids gal/1000 gal
J478 Polymers Water Water-base fluids (oil- Matrix, microfissure and
(starch) dispersible) macrofissure control, 25
to 50 lbm/1000 gal
S100 Silica Inert Water-base or oil-base Macrofissure control, 10
fluids to 35 lbm/1000 gal
U51 Diesel Oil Solvent Water-base fluids Matrix control, 5% of
total fluid volume

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix E - Fluid Loss Dowell
Page 18 of 19

7.1 Fluid-Loss Additive Sizing


Fluid-loss additives must be carefully selected to avoid invasion or plugging
problems. Particles that invade are those where the particle diameter is less than
one-sixth of the pore-throat size. Particles that plug are those where the particle
diameter is greater than one-half of the pore-throat size. Particles that bridge are
those where the particle diameter is greater than one-sixth of the pore-throat size but
less than one-half of the pore-throat size.
Particle-size selection may be determined using Fig. 7 if the pore-throat size is
known or Fig. 8 if the approximate permeability is known.

Fig. 7. Bridging particle size versus pore-throat diameter.

DOWELL CONFIDENTIAL
Section 1500
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix E - Fluid Loss
Page 19 of 19

Fig. 8. Bridging particle size versus approximate permeability.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 1 of 15

APPENDIX F - EQUIPMENT

1 Introductory Summary............................................................................................................. 2

2 Mixing and Blending Equipment ............................................................................................ 4


2.1 PCM Precision Continuous Mixer ........................................................................................ 4
2.2 POD Blender ........................................................................................................................ 5
2.2.1 POD II Blender............................................................................................................ 7

3 Pumping Equipment ................................................................................................................ 9


3.1 Pump Application Guidelines ............................................................................................... 9
3.1.1 Fracturing Fluid Viscosity............................................................................................ 9
3.1.2 Slurried Fluids Containing Proppant ........................................................................... 9
3.1.3 Proppant Concentration and Low Pump Speeds...................................................... 10
3.1.4 High Vapor Pressure Fracturing Fluids..................................................................... 10
3.1.5 Volumetric Efficiency ................................................................................................ 11
3.2 Nitrogen ............................................................................................................................. 11
3.3 Carbon Dioxide .................................................................................................................. 12
3.4 Pressure Multipliers ........................................................................................................... 12

4 Treating Equipment ............................................................................................................... 12

5 Sensors................................................................................................................................... 13

6 Computing and Monitoring Equipment ............................................................................... 13


6.1 PACR Pumping, Acidizing, Cementing Recorder............................................................... 13
6.2 PPR Pumping Parameter Recorder ................................................................................... 13
6.3 FCS Computer System ...................................................................................................... 13
6.4 PAC Portable Acquisition Computer .................................................................................. 13
6.5 JMU Job Management Unit ............................................................................................... 14

7 Tools ....................................................................................................................................... 15

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 2 of 15

8 Support Equipment ................................................................................................................15


FIGURES
Fig. 1. Equipment positioning for a fracturing treatment (typical). ................................................3
Fig. 2. The RampFRAC Service. ..................................................................................................5
Fig. 3. The stairstep method of proppant addition. .......................................................................6

1 Introductory Summary
Precision operation and dependability are two key factors in fracturing treatments.
They are necessary to ensure that the treatment is executed as designed or can be
modified in a controlled manner as dictated by well response.
Several types of equipment are necessary to successfully perform a fracturing
treatment. These are:
• mixing and blending equipment
• pumping equipment
• treating equipment
• sensors
• computing and monitoring equipment
• tools
• support equipment.
Fig. 1 (not to scale) illustrates the equipment positioning for a fracturing treatment on
a land location. The type and quantity of equipment is dependent on many
variables. Fig. 1 is for illustrative purposes only.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 3 of 15

Fig. 1. Equipment positioning for a fracturing treatment (typical).

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 4 of 15

2 Mixing and Blending Equipment


Fracturing fluids, liquid- and dry-additives, and proppants are combined, thoroughly
mixed, and discharged using mixing and blending equipment.

2.1 PCM Precision Continuous Mixer


The PCM* precision continuous mixer is a pumping and blending system that allows
continuous mixing of fracturing fluid. Water-base fracturing fluids with equivalent
polymer concentrations of 10 to 60 lbm/1000 gal can be continuously mixed and
discharged at a maximum of 70 bbl/min. Equipment consists of
• a 1500-gal liquid polymer-slurry tank (1500 gallons of polymer slurry will make
160,000 gal of WF40)
• a 10,000-gal, six-compartment reactor (for polymer hydration)
• a potassium chloride (KCl) additive system (rates up to 12 bbl/min)
• three liquid-additive systems with a storage capacity of 345 gal each (rates up to
15 gal/min)
• one slurry-additive system with a storage capacity of 345 gal (rates up to
15 gal/min).
All systems are redundant. All additive rates may be automatically or manually
adjusted. The values for fluid and additive rates, pH, temperature, conductivity and
viscosity are monitored and output to the TCV* treatment control vehicle or JMU job
management unit.
Limitations of Application
The PCM mixer has open-topped tanks containing agitators and cannot be used for
mixing oil-base fluids. An explosive oil-base cloud formed over the tanks could
easily result in a huge location fire.
Acid-base fluids should not be introduced into the PCM mixer. The PCM mixer is
manufactured with numerous aluminum components (for example, impellers, shafts,
and brackets) that are incompatible with acid-base fluids.
References
General specifications for the PCM mixer are provided in the Equipment Catalog.
Detailed PCM mixer information is provided in the SBF211 Operators Manual.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 5 of 15

2.2 POD Blender


The POD∗ programmable optimum density blender uses process-control computers
to meter the precise proppant-to-fluid ratios throughout the treatment, strictly
adhering to the treatment design parameters. This precision blending capability is
ideal for the RampFRAC* service, considered ideal for optimum proppant placement
(Fig. 2).

Fig. 2. The RampFRAC Service.


THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY
The POD blender can accurately mix and meter proppant, dry additives, liquid
additives and fracturing fluid together at a specified density in the pre-programmed

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 6 of 15

automatic mode. Equally precise metering can be handled from the command
console (truck only) or from the remote-control panel.
Response time is minimized because signals from the integral rapid-response
densitometer are routed directly to the microprocessor controlling proppant addition.
Normal response time to a change in proppant concentration is less than 10 sec,
regardless of pump rate. This immediate response time permits accurate and quick
proppant addition necessary for a stairstep treatment design (Fig. 3).

Fig. 3. The stairstep method of proppant addition.


THIS FIGURE IS FOR ILLUSTRATIVE PURPOSES ONLY
The proppant concentration can be precisely and safely controlled up to 22 PPA
(sand) and 32 PPA (high-strength proppant). Overall proppant/fluid ratio is
constantly monitored and controlled in a range of ± 0.5%.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 7 of 15

Additionally, the central processing unit maintains discharge pressure at a constant


value to ensure adequate net positive suction pressure at the high-pressure,
reciprocating-pump suction manifolds.
The POD blender can be truck-mounted or skid mounted. The truck-mounted,
remote-controlled POD blender is capable of blending and pumping slurry at
70 bbl/min and 100 psi. Equipment consists of
• two independent vortex mixers
• microprocessor-controlled integral gates to meter the proppant
• two microprocessor-controlled liquid-additive pumps (0.5 to 6 gal/min each)
• one microprocessor-controlled liquid-additive pump (0.5 to 10 gal/min)
• two microprocessor-controlled dry-additive systems (0.009 to 0.640 ft3/min each)
• a tank for liquid-additive storage.
The skid-mounted, remote-controlled POD blender is capable of blending and
pumping slurry at 35 bbl/min and 100 psi. Equipment consists of
• a vortex mixer
• a microprocessor-controlled integral gate to meter the propping agent
• a single-piece skid with a crashframe.
References
General specifications for the POD blenders are provided in the Equipment Catalog.
Detailed truck-mounted POD blender information is provided in the SBT611
Operators Manual. Detailed skid-mounted POD blender information is provided in
the SBS611 Operators Manual.

2.2.1 POD II Blender


Like the POD blender, the POD II blender is equipped with two independent vortex
mixers with microprocessor-controlled integral gates to meter the proppant.
Additionally, the POD II blender has the following improvements:
• Communication  The POD II blender will communicate with a PPR* pumping
parameter recorder or a PAC* portable acquisition computer.
• Suction and discharge process flowmeters  Flow rate information is
communicated side-to-side and to the data acquisition system. Pump rate and
volume calculation no longer depend on high-pressure pump efficiency. With the
flow rate known, additives can be set to run in the concentration mode. Since
both sides know total flow, additive systems on either side can meter according to
the total unit flow.

* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 8 of 15

• Liquid-additive system  The liquid-additive pumps are stainless-steel pumps.


The pump-rate range is 1 to 35 gal/min and the pumps can pump slurries.
Nonintrusive, coriolis flowmeters are used to measure flow rate and are much
more accurate than turbine flowmeters.
The liquid-additive system has a great deal of functional redundancy, including
calibration tanks similar to the tanks on the PCM mixer and prejob calibration of
tachometer rates to the flowmeters.
The POD II blender can meter liquid additives based on fluid density rather than
fluid rate. This means that as the liquid rate decreases (the proppant
concentration increases), the liquid-additive rates ramp-down with the fluid rate
(RampADD* Service). This provides a consistent fluid quality instead of a
stepwise quality. The FracCAT∗ Service system-software sends concentration
set-points and can use this feature. The PPR software sends rate set-points and
cannot use this feature.
• Dry-additive system  The POD II blender can meter dry additives based on fluid
density rather than fluid rate. This means that as the liquid rate decreases (the
proppant concentration increases), the dry-additive rates ramp-down with the
fluid rate. Dry-additive rates are established by entering the desired screw factor
directly into the control panel as opposed to running on speed (rpm). The
concentration can also be ramped using the FracCAT software.
• Standard remote display (SRD)  The POD II blender has incorporated all
functions into the SRD. This is the same SRD used with the PAC computer, but
configured differently for the blender. The operator switches screens to interact
with different parts of the software such as self-diagnostics, maintenance tracking
or total volumes pumped. The SRD is programmable.
• Density measuremen t Density calculation routines are incorporated in the
POD II blender software. The densitometers are calibrated and zeroed using the
control panel.
• Density control  Ramps can now start at zero density. If a densitometer fails,
the density is calculated using the flowmeters. If a flowmeter fails, old methods of
density control automatically take over.
• Self-diagnostics  Self-diagnostics provide real-time information during the job if
the software suspects equipment problems, allowing the operator to take
corrective action. Before the job, the unit will inform the operator if calibrations
(for example, densitometer calibration) have changed from previous values,
indicating possible problems. Material quantities (for example, total proppant
pumped) are recorded to aid in mechanical maintenance such as slinger wear. A
unit “health check” can be performed using a built-in self-test procedure.
Mechanical and electrical components (including sensors) are exercised while
the software searches for anomalies.

∗ Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 9 of 15

3 Pumping Equipment
High-pressure pumping equipment provides the necessary horsepower to create and
propagate the fracture.
Fracturing fluid is pumped with reciprocating pumps. A reciprocating pump is a
mechanical device used to impart a pulsating, dynamic flow to a liquid and consists
of one or more single- or double-acting positive displacement elements (pistons or
plungers). The elements in the fluid end are driven in harmonic motion by a slider
crank mechanism. The liquid flow generated by this reciprocating motion is directed
from the pump inlet (suction) to the pump outlet (discharge) by the selective
operation of self-acting check valves located at the inlet and outlet of each
displacement element.
The reciprocating pumps used by Dowell are single-acting triplex (three cylinder) and
quintiplex (five cylinder) pumps. Diesel and turbine engines are used to provide
power to operate the pumps.

3.1 Pump Application Guidelines

3.1.1 Fracturing Fluid Viscosity


The limited amount of available literature in pump textbooks and handbooks on
reciprocating pump performance invariably limits the discussion to water (viscosity of
1 cp). Unfortunately, low-viscosity fracturing fluids are rarely used. In general, very
limited testing indicates that the viscosities typically encountered in linear fracturing
fluids do not have a significant effect on pump performance.

3.1.2 Slurried Fluids Containing Proppant


Limited testing with Dowell pumps has been performed to determine the effect of
proppant concentration on pump performance. Testing has been limited to water-
base fracturing fluids containing sand as a proppant. The testing has revealed that
valve dynamics can be extremely sensitive to proppant concentration.
The poor pump performance manifests itself as rough running or line shake. As
such, it can easily be mis-diagnosed as cavitation; however, poor pump performance
at relatively high proppant rates and pump speeds (when suction pressure is
adequate) may be caused by delayed closure or “valve float.”
Modification of pump valves and springs has improved the slurry handling
capabilities of some pumps. Testing with the QL and QI quintiplex pumps has
indicated a maximum proppant concentration of 13 PPA at 20 bbl/min. The
maximum proppant concentration is reduced to 3 PPA at 29 bbl/min. Limitations
exist for all Dowell pumps; however additional testing has not been performed.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 10 of 15

3.1.3 Proppant Concentration and Low Pump Speeds


The addition of a recirculation line to suction manifolds on stimulation units pumping
high proppant concentrations at low pump speeds is customary. A recirculation line
increases the mean fluid velocity through the suction system by the recirculation of a
portion of the flow back to the blender. This decreases the tendency of the proppant
to settle in the pump suction header. If sufficient proppant does settle in the suction
header, the belief is that slugs of proppant can be ingested by the pump, totally
blocking one or more cylinders. Proppant settling will be worst at the cylinder
farthest from the suction header inlet.
Although the velocity profiles in the suction line and header are known, the actual
mechanism of proppant packing in the pump is not well understood. The proppant
settling, to some extent, is self-compensating. As proppant settles, the conduit cross
sectional area is reduced which increases mean velocities and reduces the tendency
of the proppant to settle.
Field practice varies widely as to determination of when to use a recirculation line.
Factors to consider are
• pump speed
• mean suction velocity
• fluid properties
• proppant type
• proppant concentration.
The practice is most prevalent when proppant concentrations exceed 10 lbm
proppant added and pump speed is less than 150 rpm.

3.1.4 High Vapor Pressure Fracturing Fluids


High vapor pressure fracturing fluids reduce the available net positive suction head
to the pump. Parameters to consider are
• Water-base fluids at high temperature. The vapor pressure of water rapidly rises
when the fluid temperature exceeds 140°F (60°C).
• Inflammable or combustible fluids. A requirement of Dowell Location Safety
Standard 5 is to determine the vapor pressure of fracturing fluids containing oil,
condensate, alcohol or solvent(s) prior to use. Generally, the vapor pressure of
these fluids will not significantly reduce the net positive suction head available.
• Carbon dioxide (CO2). The vapor pressure of CO2 at -109°F (-43°C) is one
atmosphere. Carbon dioxide will revert to the vapor phase in normal pumping
conditions. Therefore, CO2 is pumped at high (suction) pressure (300 to 350
PSIG) to ensure that no vapor phase exists or can easily form. A special CO2
suction header is installed on the pump and must be used when pumping CO2.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 11 of 15

3.1.5 Volumetric Efficiency


Reciprocating pumps never achieve theoretical flow rates for two reasons.
1. Operational inefficiencies (cavitation and valve float) in the pump. When
operating under conditions of cavitation or valve float, discrepancies will always
arise between mechanical barrel counters and actual flows.
2. Fracturing fluid compressibility. Fluid compressibility is an intrinsic property of
fluid and is unavoidable.
Volumetric efficiency is the ratio of the actual flow rate to the theoretical flow rate.
Most Dowell mechanical barrel counters are calibrated assuming a volumetric
efficiency of 97%. Volumetric efficiencies are usually not a concern in the majority of
Dowell services.
References
Additional information is provided in the Reciprocating Pump Application Manual and
Dowell Location Safety Standards. Fluid end ratings and constants are provided in
the Treating Equipment Manual.

3.2 Nitrogen
Nitrogen (N2) is delivered to the wellsite in a liquid state at -320°F (-196°C) and must
be vaporized before being pumped. A centrifugal pump feeds the liquid nitrogen to a
high pressure reciprocating pump. The reciprocating pump, in turn, pumps the liquid
nitrogen through a vaporizer (heat exchanger). The vaporizer converts the liquid
nitrogen to a gas and heats it to 80 to 100°F (27 to 38°C). The vaporizer may be
either a diesel-fired unit or a (flameless) water brake. Diesel-fired equipment may be
used where local regulations permit operation. Flameless equipment is used
offshore.
Another type of nitrogen equipment is the Pressure Swing Adsorption (PSA) unit.
This equipment allows onsite generation of high-pressure nitrogen gas drawn from
ambient air. This is advantageous when liquid nitrogen is difficult to obtain or in
remote areas. A disadvantage of PSA equipment is the low discharge rate
(approximately 650 scf/M at 5000 psi).
Nitrogen pumping equipment may be either truck- or skid mounted.
References
General descriptions and specifications for nitrogen pumping equipment are
provided in the Equipment Catalog. Safety information is provided in the Dowell
Location Safety Standards.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 12 of 15

3.3 Carbon Dioxide


Carbon dioxide (CO2) is delivered to the wellsite in a liquid state at 0°F (-18°C).
Unlike nitrogen, carbon dioxide is not converted to a gas and preheated before being
pumped. A centrifugal pump (CO2 booster pump) feeds the liquid carbon dioxide at
300 to 350 psi to conventional reciprocating pumps equipped with a special suction
manifold.
Carbon dioxide pumping equipment may be either truck- or skid mounted.
References
Additional information is provided in the Dowell Location Safety Standards.

3.4 Pressure Multipliers


Pressure multipliers are intended to operate at treating pressures from 10,000 to
20,000 psi for long periods of time (greater than two hours). The pressure multiplier
is equipped with a duplex pump and is powered with a turbine engine. The duplex
pumps have a stroke of 70 in. or more. This allows fewer pump strokes and valve
cycles per volume of fluid pumped and greatly prolongs pump life.

4 Treating Equipment
Treating equipment is the equipment, other than trucks or skids that is used to
execute a fracturing treatment. Treating equipment includes
• swivel joints
• valves
• treating adapters
• pipes and loops
• hose and fittings
• ball injectors
• fracturing heads
• fluid ends
• tree savers (wellhead protectors).
Dispersers, quick disconnects and other miscellaneous equipment are also treating
equipment.
References
Comprehensive information for the maintenance and operation of treating equipment
is provided in the Treating Equipment Manual. Safety and use information is
provided in the Dowell Location Safety Standards.

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 13 of 15

5 Sensors
A sensor provides the result, or, “measured data” in the form of an analog signal.
Examples of measured data are the wellhead pressure, pump rate and slurry
density. Measured data are used to provide “calculated data.” Examples of
calculated data are bottomhole pressure, volume and proppant concentration.
References
Comprehensive information for the maintenance and operation of sensors is
provided in the Sensor Engineering Manual and the Sensor Verification Manual.
Additional information is provided in the various operation and technical manuals for
sensors.

6 Computing and Monitoring Equipment


The fracturing treatment effectiveness is greatly enhanced by using computing and
monitoring equipment to process and evaluate treatment parameters before, during
and after the fracturing treatment.

6.1 PACR Pumping, Acidizing, Cementing Recorder


The PACR* pumping, acidizing, cementing recorder is a portable data acquisition
system that monitors and records sensor signals at the wellsite during pumping
operations.

6.2 PPR Pumping Parameter Recorder


The PPR pumping parameter recorder is a portable stimulation process monitoring
and control system. The PPR allows the mobile deployment of microprocessor-base
equipment used to monitor, evaluate and control well treatments.

6.3 FCS Computer System


The FCS computer system is based on the MicroVAX+ hardware and allows the
mobile deployment of a computer to support the CADE* Computer-Aided Design and
Evaluation software.

6.4 PAC Portable Acquisition Computer


The PAC portable acquisition computer is a standalone acquisition system with a
Standard Remote Display (SRD). It is primarily used as a front-end acquisition
system that supplies job-related data to Dowell wellsite control, reporting and
analysis systems (Wellsite Reporting System [WRS] software or FracCAT Service
system-software. The PAC computer is a ruggedized unit that acquires data from all
Dowell equipment and certified sensors. It can easily connect with, and monitor,

* Mark of Schlumberger
+ Trademark of Digital Equipment Corporation

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
May 1998 Schlumberger
Appendix F - Equipment Dowell
Page 14 of 15

additional sensors that utilize standardized petroleum industry output signals. The
SRD allows an operator to quickly program, calibrate and operate the acquisition
computer while monitoring job parameters in real time.
The PAC computer can send up to 40 parameters to a Dowell Wellsite Reporting
System (WRS) or to a client's personal computer. When used in conjunction with the
FracCAT Service, the PAC computer completes the communications loop between
Dowell process-control equipment and the analysis computers. Acting like a bridge,
the PAC computer allows commands from the analysis computer to reach the
process-control equipment. This equipment carries out the designated commands,
then reports back to the analysis computer (via the PAC computer).
The PAC computer has a NEMA-4X (IP65) water and dust protection rating and
battery backup for internal RAM memory. The operating temperature ranges from -
20 to 70°C (-4 to 158°F).
PAC computer features are
• supports oilfield, Canadian metric and French metric units
• communicates with Dowell process-control equipment (POD Blender, VIP* Mixer
and CLAS* system)
• an external mass-storage drive which uses removable memory-card technology
is optional
• no moving parts in either the acquisition computer or the optional external
storage drive
• accepts up to 60 inputs (sensors)
• compatible with existing Dowell sensors (standard and custom)
• pressure and density are updated at a 2-Hz rate
• can accept bottomhole pressure and bottomhole temperature measurements
from all major wireline companies
• user-selectable record rate from 0.5 to 10 sec
• recorded data can be played back
• postjob corrections can be applied and played back if a sensor's original setup
was not correct
• self-guiding, troubleshooting program is built into the system
• the PAC computer is portable, but may be permanently mounted.

6.5 JMU Job Management Unit


The JMU job management unit allows the mobile, onsite deployment of the FracCAT
Service. The client can use the FracCAT Service system-software in the JMU.
Externally, the JMU resembles the TCV treatment control vehicle. Internally, the JMU
is divided into management and operator compartments. The management
* Mark of Schlumberger

DOWELL CONFIDENTIAL
Section 1600
FRACTURING ENGINEERING MANUAL
Schlumberger May 1998
Dowell
Appendix F - Equipment
Page 15 of 15

compartment, or client/service supervisor area, houses the treatment control


hardware and software. Five people can be comfortably seated, and a total of
10 people may be accommodated in the management area to observe and use the
FracCAT software. The operator compartment is arranged so that equipment
operators can maintain eye and voice contact with the Dowell service supervisor.
Six remote-control panels (expandable to nine) can be placed into the operator
compartment. An uniterruptible power supply assures continued operation in the
unlikely event that the direct-drive AC generator fails.
References
General specifications for computing and monitoring equipment are provided in the
Equipment Catalog. Additional information is provided in the Sensor Verification
Manual, Sensor Engineering Manual, FracCADE User Manual, Wellsite Reporting
System User's Manual, and the various operation and technical manuals for
computing and monitoring equipment.

7 Tools
Packers and bridge plugs are mechanical devices used to protect casing strings and
divert fracturing fluid. Methodology is provided in Diverting Techniques. Additional
information is provided in the Downhole Tools Hydraulic Manual and the Retrievable
Tools Technical Manual. General specifications for tools are provided in the
Equipment Catalog.

8 Support Equipment
Support equipment is the equipment used for handling materials. Support
equipment includes
• liquid-transport trailers
• transfer pumps
• manifolds
• dry-material transports
• dump trucks
• proppant feeders
• proppant conveyers
• utility trucks
• portable tanks.
General descriptions and specifications for support equipment are provided in the
Equipment Catalog.

DOWELL CONFIDENTIAL

You might also like