You are on page 1of 162

-_-_._. ......_.-._._._---- ----------=---- ..

_----_.

Hydraulic Fracture Mechanics


-"'-"-------"'-:---------------------~

Hydraulic Fracture Mechanics

Peter Valko and Michael J. Economides


TexasA & M University,
College Station,
USA

JOHN WILEY & SONS


Chichester. New York. Brisbane. Toronto. Singapore
Copyright 1995 by John Wiley & Sons Ltd,
Baffins Lane, Chichester,
West Sussex P019 IUD, England

National 01243 779777


International (+44) 1243 779777

Reprinted October 1996, May 1997

All rights reserved.


CONTENTS
No part o~ this book may be reproduced by any means,
or transmitted, or translated into a machine language
without the written permission of the publisher.

Other WIley Editorial Offices

John Wiley & Sons, Inc., 605 Third Avenue, Preface xi


New York, NY 101580012, USA
list of Notation xiii
Jagaranda Wiley Ltd, 33 Park Road, Milian,
Queensland 4064, Australia
1 Hydraulically Induced Fractures in the Petroleum
John Wiley & Sons (Canada) Ltd, 22 Worcester Road, and Related Industries 1
Rexdale, Ontario M9W Ill, Canada
1.1 Fractures in Well Stimulation 1
John Wiley & Sons (SEA) Pte ltd, 37 Jalan Pemimpin #0504, 1.2 Fluid Flow Through Porous Media 2
Block B, Union Industrial Building, Singapore 2057 1.2.1 The Near-well Zone 4
1.3 Flow from a Fractured Well 5
1.4 Hydraulic Fracture Design 7
1.5 Treatment Execution 11
1.5.1 Fracturing Fluids 11
1.5.2 Proppants 13
1.6 Data Acquisition and Evaluation for Hydraulic Fracturing 14
1.6.1 Well Log Measurements 14
1.6.2 Core Measurements 15
1.6.3 Well Testing 15
1.7 Mechanics in Hydraulic Fracturing 15
References 16

2 Linear Elasticity, Fracture Shapes and


Induced Stresses . 19
2.1 Force and Deformation 19
2.1.1 Stress 19
2.1.2 Strain 21
British Library Cataloguing in Publication Data 2.2 Material Properties 23
2.2.1 Linear Elastic Material 23
A catalogue record for this book is available from the British library 2.2.2 Material Behavior Beyond Perfect Elasticity 26
2.3 Plane Elasticity 27
ISBN 0 471 956643 2.3.1 Plane Stress 27
2.3.2 Stresses Relative to an Oblique Line
Typeset in 10~/12i TImes by Laser Words, Madras, India (Force Balance I) 28
Printed and bound in Great Britain by Bookcraft (Bath) Ltd 2.3.3 Equilibrium Relations (Force Balance II) 30
2.3.4 Plane Strain 30
2.3.5 Boundary Conditions 32
-_--- ...._._-------

vi Contents Contents vii

2.4 Pressurized Crack 32 5.2 Slot Flow 105


2.4.1 Solution of the Line Crack Problem 32 5.2.1 Derivation of the Basic Relations 105
2.4.2 Constant Pressure 34 5.2.2 Equivalent Newtonian Viscosity 111
2.4.3 Polynomial Pressure Distribution 35 5.3 Flow in Circular Tube 112
2.4.4 "Zipper" Cracks 37 5.3.1 Basic Relations 112
2.4.5 "Zipper" Crack with Polynomial Pressure Distribution 40 5.3.2 Flow Curve 115
2.5 Stress Concentration and Stress Intensity Factor 41 5.3.3 Equivalent Newtonian Viscosity for Tube Flow 119
2.5.1 Stress Intensity Factor, Symmetric Loading 42 5.4 Flow in Other Cross Sections 122
2.5.2 Stress Intensity Factor, non-symmetric Loading 43 5.4.1 Flow in Annulus 122
2.6 Fracture Shape in the Presence of Far-field Stress. 5.4.2 Flow in Elliptic Cross Section 123
The Concept of Net Pressure 43 5.4.3 Limiting Ellipsoid Cross Section 124
2.7 Circular Crack 45 References 128
2.8 Volume and Strain Energy 47
2.9 Computational Methods 49
References 50 6 Non-laminar Flow and Solids Transport 131
6.1 Non-laminar Flow 131
6.1.1 Newtonian Fluid 131
3 Stresses in Formations 53 6.1.2 General Fluid 132
3.1 Basic Concepts 6.1.3 Drag Reduction 134
53
3.2 Stresses at Depth 6.1.4 Turbulent Flow in Other Geometries 137
55 138
3.3 Near-wellbore Stresses 59 6.2 Solids Transport
3.4 6.2.1 Settling of an Individual Sphere 139
Stress Concentrations for an Arbitrarily Oriented Well 63
3.5 6.2.2 Effect of Shear Rate Induced by Flow 141
Vertical Well Breakdown Pressure 65
6.2.3 Effect of Slurry Concentration 142
3.6 Breakdown Pressure for an Arbitrarily Oriented Well 66 6.2.4 Wall Effects 143
3.7 Limiting Case: Horizontal Well 69 6.2.5 Agglomeration Effects 145
3.7.1 Arbitrarily Oriented Horizontal Well 70 References 145
3.8 Permeability and Stress 71
3.8.1 Stress-sensitive Permeability 72
3.9 Measurement of Stresses 73 147
7 Advanced Topics of Rheology and Fluid Mechanics
3.9.1 Small Interval Fracture Injection Tests 74
3.9.2 Acoustic Measurements 75 7.1 Foam Rheology 147
3.9.3 Determination of the Closure Pressure 76 7.1.1 Quality Based Correlations 148
3.9.4 Core Stress Measurements 77 7.1.2 Volume Equalized Constitutive Equations 148
3.9.5 Critique and Applicability of Techniques 79 7.1.3 Volume Equalized Power Law 151
References 7.1.4 Turbulent Flow of Foam 152
80
7.2 Accounting for Mechanical Energy 153
7.2.1 Basic Concepts 153
4 Fracture Geometry 83 7.2.2 Incompressible Flow 154
7.2.3 Foam Flow 154
4.1 The Perkins and Kern and Khristianovich and 156
7.3 Rheometry
Zheltov Geometries 83 156
7.3.1 Pipe Viscometry
4.1.1 The Consequences of the Plane Strain Assumption 86 7.3.2 Slip Correction 157
4.2 Fracture Initiation vs. Propagation Direction 88 162
References
4.2.1 Fractures in Horizontal Wells 90
4.3 Fracture Profiles in Multi-layered Formations 92
References 95 8 Material Balance 165
8.1 The Conservation of Mass and Its Relation to
5 Rheology and Laminar Flow Fracture Dimensions 165
97 169
8.2 Fluid Leakoff and Spurt Loss as Material Properties
5.1 Basic Concepts 97 8.2.1 Carter Equation I 169
5.1.1 Material Behavior and Constitutive Equations 98 8.2.2 Formal Material Balance. The Opening Time
5.1.2 Force Balance 103 Distribution Factor 171
__
........ _
..... --------

viii Contents Contents ix

8.3 The Constant Width Approximation (Carter Equation II) 172 10.3 Retarded Fracture Propagation 245
8.4 The Power Law Approximation to Surface Growth 174 10.3.1 Fluid Lag 245
8.4.1 The Consequences of the Power Law Assumption 174 10.3.2 TIp Dilatancy 245
8.4.2 The Combination of the Power Law Assumption 10.3.3 Apparent Fracture Toughness 246
178 10.3.4 Process Zone Concept 246
with Interpolation
10.3.5 The Reopening Paradox 247
8.5 Numerical Material Balance 179
10.4 Continuum Damage Mechanics in Hydraulic Fracturing 247
8.6 Differential Material Balance 181
10.4.1 TIp Propagation Velocity from COM 247
8.7 Leakoff as Flow in the Porous Medium 183
10.4.2 CDM-NK Model 249
8.7.1 Filter-cake Pressure Drop 184
10.4.3 CDM-PKN Design Model 252
8.7.2 Pressure Drop in the Reservoir 185
10.5 Pressure Decline Analysis and TIp Retardation 256
8.7.3 Leakoff Rate from Combining the Resistances 187
10.5.1 Resolving Contradictions with Continuum
(Ehlig-Economides et al. [6])
Damage Mechanics 258
References 187
References 263

9 Coupling of Elasticity, Flow and Material Balance 189


9.1 Width Equations of the Early 20 Models 189 11 Fracture Height Growth (3~ and P-3D Geometries) 267
9.1.1 Perkins-Kern Width Equation 189 11.1 Equilibrium Fracture Height 269
9.1.2 Geertsma-de Klerk Width Equation 192
11.1.1 Reverse Application of the Net-pressure Concept 269
9.1.3 Radial Width Equation 195 11.1.2 Different Systems of Notation 270
9.2 Algebraic (20) Models as Used in Design 196 11.1.3 Basic Equations 272
9.2.1 PKN-C 196 11.1.4 The Effect of Hydrostatic Pressure 276
9.2.2 KGD-C 199
11.2 Three-dimensional Models 278
9.2.3 PKN-N and KGD-N 200
11.2.1 Surface Integral Method 279
9.2.4 PKN-a and KGD-a 201
11.2.2 The Stress Intensity Factor Paradox 281
9.2.5 Radial Model 202 11.3 Pseudo-three-dimensional Models 283
9.2.6 Non-Newtonian Behavior 202
References 284
9.3 Numerical Material Balance (NMB) with Width Growth 204
9.4 Differential 20 Models 205
9.4.1 Nordgren Equation 206
9.4.2 Differential Horizontal Plane Strain Model 209 Appendix: Comparison Study of Hydraulic Fracturing
9.5 Models With Detailed Leakoff Description 210 Models: Input Data and Results 287
9.6 Pressure Decline Analysis 211 References 294
9.6.1 Nolte's Pressure Decline Analysis
(Power Law Assumption) 212 Index 295
9.6.2 The No-spurt-Ioss Assumption
(Shlyapobersky method) 217
9.6.3 Material Balance and Propagation Pressure
Estimates of the Spurt Loss 218
9.6.4 Resolving Contradictions 227
9.6.5 Pressure Decline Analysis With Detailed Leakoff
Description (Mayerhofer et al. Technique) 230
References 232

10 Fracture Propagation 235


10.1 Fracture Mechanics 237
10.1.1 Griffith's Analysis of Crack Stability 238
10.1.2 Mott's Theory for the Rate of Crack Growth 241
10.2 Classical Crack Propagation Criterion for
Hydraulic Fracturing 242
10.2.1 Fracture Toughness Criterion 242
10.2.2 The Injection Rate Dependence Paradox 243
---------------------------------~-~- ..~.----~-
.....--._--

PREFACE

This book addresses the theoretical background of one of the most widespread activi-
ties in hydrocarbon wells, that of hydraulic fracturing. It provides a treatment of basic
phenomena including elasticity, stress distribution, fluid flow, and the dynamics of
the rupture process from the point of view of the influence of those phenomena
on the created fracture. Currently used design and analysis techniques are derived
and improved using a comprehensive and unified approach. Numerical ~xamples are
elaborated to illustrate important concepts.
The material grew out of university and industrial courses that have been taught
at Mining University of Leoben, Texas A&M University and several locations
throughout the world. During these courses we have recognized that currently
available monographs, often written bya great number of co-authors reflect diverse
views, systems of notations and units. One of our main goals was to establish a
common language that eases the way workers in the field can get acquainted with
the material and experts of different background can communicate with each other.
Our gratitude goes to our coworkers and students who have contributed
a great deal to the final form of the book. The list below is far from
complete: T. Brugger, H. Buchsteiner, Zhongming Chen, C. Enzendorfer Yong
Fan, T.P. Frick, M.J. Mayerhofer, H. Mosser, R. Oligney, W. Prassl, M. Prohaska,
C.R. Rom, R.E. Schmid, 1. Smith, R. Seiler, W. Winkler, and M. Zettl.
Anybody interested in hydraulic fracturing is bound to be influenced by the
pioneering work down in the fifties and sixties. The authors of this book have had
the privilege to enjoy discussions with the developers of the first and compelling
models including YP. Zheltov and T.K. Perkins, If. The hand-written remarks of
I. Geertsma are saved with particular honor. While the views expressed on these
pages have also been influenced by personalities such as M.P. Cleary, S.A. Holditch,
M.K. Hubbert, K.G. Nolte, 1. Shlyapobersky and N.R. Warpinski, we take full
. responsibility for the content and format of presentation.
We would like to express our gratitude to organizations for permitting us to
reproduce some of the figures and tables in the text: Society of Petroleum Engi-
neers (Figure 1-1, Figure 1-2, Figure 3-12, Figure 3-14, Figure 4-7, Figure 11-4,
and Tables AI-AS); American Institute of Physics, (Figure 7-5, Figure 7-6, and
Figure 7-7).

The Authors
-------- .. ~---- .-.~-------

LIST OF NOTATION

7
A area, m-
A' Reidenbach et al. constant coefficient, kgQ,77 m-L54 . s-o 77 (Ch, 7)
Ac cross sectional area for flow, m2
AD dimensionless fracture surface area
A"Afe fracture surface area at end of pumping, m2 (Ch, 8,9)
Afb fracture area per unit of bulk volume, m" (Ch.8)
AfD dimensionless fracture network area (Ch. 8)
Afma fracture area per unit of matrix volume, m" (Ch. 8)
AL fracture surface area exposed to leakoff, m2 (Ch. 8)
An,Aj fracture surface areas at different time instants, m2
Bx,By conductivity/porosity factor, m2/3 (Ch. 3)
C Kachanov parameter of damage accumulation rate, Pa-~s-l (Ch. 10)
CD dimensionless Kachanov parameter (Ch. 10)
CD drag coefficient, dimensionless (Ch. 7)
CD,., modified drag coefficient, dimensionless (Ch. 7)
Cij element of the linear elasticity coefficient matrix, Pa
CL leakoff coefficient, m.s-I/2
CL,Q coefficient of pressure dependent leakoff, rn- S-I/2. Pa-1 (Ch, 11)
D diameter, well diameter, m
D damage variable, dimensionless (Ch. 10)
Dv dissipation rate, J . S-1
D1,D2 annulus smaller and larger diameter, m (Ch, 5)
E Young's modulus, Pa
Ee effective elastic modulus, Pa (Ch, 11)
E' plane strain modulus, Pa
Ei exponential integral
E(k),E(m) complete elliptic integral of the second kind
F force, N
Fc dimensionless correction factor for proppant settling
FCD dimensionless fracture conductivity (Ch. 1)
F CD,opt dimensionless fracture conductivity, optimal (Ch. 1)
G shear modulus, Fa
G(~) auxiliary function for circular crack
H formation depth, m (Ch. 1,2)
[CDM CDM width factor, dimensionless (Ch. 10)
xiv List of Notation List of Notation xv

distortional creep compliance, Pa-1 (Ch. 3) energy, J


dilatational creep compliance, Pa-I (Ch. 3) work of inner pressure to move fracture faces apart, J
consistency index, Pa- s- kinetic energy, J
stiffness matrix, Pa . m (Ch, 11) strain energy caused by net pressure, J
K' generalized consistency index, Pa . s-" strain energy caused by far-field stress, J
K~ generalized consistency index for pipe flow, Pa- s-n compressibility factor of gas, dimensionless
K/ stress intensity factor (mode I), Pa- m1/2 a exponent (Ch. 5)
K/c critical stress intensity factor, fracture toughness, Pa- m 1/2 a coefficient, m2. (Pa '5)-1 (Ch. 7)
K1C,boUOm fracture toughness at bottom, Pa . m 1/2 auxiliary variables, dimensionless (Ch. 5)
KIC.IOp fracture toughness at top, Pa- m 1/2 auxiliary variable, m3/4 (Ch. 9)
K1." nominal stress intensity factor, Pa- m1/2 (Ch, 10) auxiliary variable, Pa . m-1/4 (Ch. 9)
Kp.VE geometry dependent volume equalized consistency index (pipe), Pa- s-n b intercept, m3 (Ch, 8)
(Ch.7) b intercept, Pa (Ch. 9,10)
Kv volume equalized consistency index, Pa . s-n (Ch, 7) b coefficient, m2 . Pa-2 S-1 (Ch, 7)
Kl auxiliary coefficient, Pa-I (Ch. 7) bl fracture width, m (Ch, 8)
K2 auxiliary coefficient, dimensionless (Ch. 7) bo, b!,b2 auxiliary variables, dimensionless (Ch. 5)
K3 auxiliary coefficient, dimensionless (Ch. 7) C half-length of two-dimensional line crack, m
K2 rate of pressure increase, Pa- S-1 (Ch, 8) cf
proportionality constant in the pressure vs. width relationship, Pa- m-1
L length, m Cf.KGD
proportionality constant in the pressure vs. width relationship (KGD),
L length of contact, m (Ch, 4) Pam-1
NDe Deborah number, dimensionless proportionality constant in the pressure vs. width relationship (PKN),
NRe Reynolds number, dimensionless Pa-rn"!
NRe.p Particle Reynolds number, dimensionless proportionality constant in the pressure vs. width relationship (radial),
NRe,w wall Reynolds number, dimensionless Pa-rn"!
N~(! generalized Reynolds number, dimensionless c, total reservoir compressibility, Pa-1
R radius of a circular crack, m (Ch. 2) Clf
total fissure compressibility, Pa-!
R distance between two points, m (Ch, 11) CI
Nordgren coefficient of dimensionless length, m
RD dimensionless filter-cake resistance (Ch, 8,9) C2 Nordgren coefficient of dimensionless time, s
Rmb estimate of fracture radius, from material balance, m (Ch. 9,10) C3
Nordgren coefficient of dimensionless width, m
Rnsp estimate of fracture radius, from no-spurt-loss, m (Ch, 9,10) C. Nordgren coefficient of dimensionless net pressure, Pa
Ro filter-cake resistance, m-! (Ch. 8,9) dp particle diameter, m (Ch. 7)
Rp, estimate of fracture radius, from propagation pressure, m (Ch, 9,10) f Fanning friction factor, dimensionless
S stress vector, Pa (Ch. 2) f. factor of wall effect, dimensionless (Ch, 7)
Sp spurt-loss coefficient, m g acceleration of gravity, m- S-2
Sp.mb estimate of spurt-loss coefficient from material balance, m g(~) auxiliary function for line crack
Sp.~, estimate of spurt-loss coefficient from propagation pressure, m g(6tD. Oi) Nolte's g-function
S" s, components of the stress vector, Pa (Ch. 2) go(a) Nolte's go-function
T formation tensile strength, Pa h reservoir thickness, m
T matrix of areal elements, m2 (Ch. 11) h productive eight of fracture, m (Ch. 1)
V volume, crack volume, m3 hf fracture height, m
Vb bulk volume, rrr' hi.", fracture height provided by the modeler, m (Ch. 8)
V, fracture volume at end of pumping, rrr' : matrix thickness, m (Ch, 8)
VI fracture volume, m3 hp perforated (target) height, m
Vi volume of injected fluid, m3 injection rate per one wing, m3 . s-!
VL leakoff volume, m3 k permeability, m2
VLe leakoff volume at end of pumping, m3 k argument of elliptic integral (Ch. 11)
V"", matrix volume, m3 (Ch. 8) k Matt's numeric factor (Ch. 10)
.-._-_._....
---._.
__-.
..
_----=- .-~ ~.-~---~-~~~------------------------

xvi List of Notation List of Notation xvii

kr permeability of proppant pack in fracture, m" qz flow rate outside the plug region, m3. S-1 (Ch. 5)
krD bulk formation permeability, m2 (Ch. 8) r radial distance, m (Ch. 1)
kma matrix permeability, m2 (Ch. 8) r distance from tip, m (Ch, 2)
ko. kl k2, k3 auxiliary variables in CDM-PKN model (Ch. 10) outer boundary radius, m (Ch, 1)
I distance, m wellbore radius, m (Ch. 1)
characteristic length of flow channel, m s skin effect, dimensionless (Ch. 1)
average distance of microcracks. m (Ch. 10) equivalent skin effect, dimensionless (Ch. 1)
dimensionless average distance of microcracks (Ch. 10) repulsive distributed force, Pa (Ch. 2)
m slope, m; . s-IIZ (Ch, 8) cohesive distributed force, Pa (Ch. 2)
m slope, Pa (Ch. 9,10) time, s
m argument of elliptic integral (Ch. 5) to characteristic time (Ch. 8)
m mass flux, kg. S-1 tc closure time, s
m(p) real gas pseudo pressure, Pa- S-1 to dimensionless time
n flow behavior index, dimensionless tDxf dimensionless time for fracture + reservoir system
n' generalized flow behavior index, dimensionless [I relaxational time constant for the distortional creep, s (Ch. 3)
number of time steps in numerical material balance method tz relaxational time constant for the dilatational creep, s (Ch. 3)
P pressure, Pa t, time at end of pumping, s
Pc closure pressure, Pa t j, In different time instants, S
Pcp pressure at the center of perforation, Pa (Ch. 11) U fluid velocity, m . S-1
PD dimensionless pressure U dislacement, m (Ch. 2,10)
p, outer boundary reservoir pressure, Pa (Ch, 1) average fluid velocity, m- S~1 (Ch. 5)
Ii average reservoir pressure, Pa (Ch. 1) compressional wave slowness, m s" (Ch. 3)
P average pressure in a GDK fracture, Pa (Ch. 9) fracture propagation rate (tip velocity), m- S-1 (Ch. 8-10)
Po' initial pressure, Pa (Ch. 1) Umax maximum fluid velocity, rn- S-1 (Ch. 5)
Pis instantaneous shut-in pressure in stress determination test, Pa (Ch. 3) Us shear wave slowness, m S-1 (Ch. 3)
Pn net pressure, Pa terminal settling velocity, m S-l (Ch, 7)
Pn,tip tip net pressure, Pa (Ch. 9) terminal settling velocity, m . S-I (Ch. 7)
Pn.w wellbore net pressure, Pa (Ch. 9,10) terminal settling velocity with wall effect, rn S-1 (Ch. 7)
pnO constant net pressure in crack, Pa (Ch. 2) slip velocity, m . S-1 (Ch. 2)
PPT fracture propagation pressure, Pa (Ch. 9,10) displacement components, m (Ch. 2)
Pres.D dimensionless reservoir pressure, Pa longitudinal wave velocity, m- S-l (Ch. 2)
P",r wellbore flowing pressure, Pa (Ch. 1) leakoff velocity, m S-1 (Ch, 8)
pw.isi wellbore instantaneous shut-in pressure, Pa (Ch. 9,10) fracture width, m (Ch. 1)
Pw.pr wellbore propagation pressure, Pa (Ch. 9,10) W width of flow channel, m (Ch. 5)
Po constant pressure in line crack, Pa (Ch. 2) average fracture width at end of pumping, m
PI coefficient of polynomial pressure distribution in line crack, Pa- m-I estimate of average width at instant of shut-in, m
(Ch. 2) width of longitudinal-to-transverse transition, m (Ch. 4)
coefficient of polynomial pressure distribution in line crack, Pa- m-z W, width of ideal transverse fracture, m (Ch. 4)
(Ch. 2) fracture width at wellbore (KGD), rn
P3 coefficient of polynomial pressure distribution in line crack, Pa . m-3 Ww.O maximum fracture width at wellbore (PKN), m
(Ch.2) Ww.O.PKN maximum fracture width at wellbore, m (PKN)
production rate, m3 . -I (Ch. 1) W~'D dimensionless fracture width at wellbore (NK)
flow rate in flow channel, tube, fracture, m3 . S-I Wo maximum width of a line crack, m (Ch, 2)
dimensionless production rate (Ch. 1) Wo maximum width of the elliptical cross section, m
dimensionless flow rate into reservoir (Ch. 8,9) WO,D dimensionless maximum fracture width at wellbore (NK)
flow rates into reservoir at different times, m3 S-1 (Ch, 8,9) W~=x.f fracture width at tip (Ch. 8-10)
plug flow rate, m3 . S-1 (Ch. 5) W average fracture width, m
----_ _-_ ..
.. .._ ..... _._._---

xviii List of Notation List of Notation xix

WGDK average fracture width from GDK width equation, m Y angle characterizing direction of horizontal well, rad (Ch. 3)
WPKN average fracture width from PKN width equation, m Y auxiliary variable in BNS equation, dimensionless (Ch. 6)
X lateral coordinate, m YCDM (CDM) geometry factor, dimensionless (Ch. 10)
Xo dimensionless lateral coordinate Y shear rate, S-1
XOj dimensionless lateral coordinate at time instant j Y. wall shear rate, S-I
xf fracture half length, m Y"",,, wall shear rate (Newtonian fluid) or nominal Newtonian wall shear rate,
Xf apparent fracture half-length, m (Ch. 1) S-I (Ch. 5)

xfO dimensionless fracture length Y. average wall shear rate, S-1 (Ch. 5)
XI' fracture half-length at end of pumping, m s thickness of a slice of material, m
X t.s- Xf.n fracture half-length at different time instants, m e strain, dimensionless
X/.m fracture length provided by the modeler, m (Ch, 8) e roughness, dimensionless (Ch. 6)
x l.max maximum estimate of fracture length, m e specific volume expansion ratio, dimensionless (Ch, 7)
Xf.mb estimate of fracture length from material balance, m en horizontal strain, dimensionless
Xf,pr estimate of fracture length from propagation pressure, m e. vertical strain, dimensionless
XJ.III- estimate of fracture length from unretarded propagation , m rJ fluid efficiency, dimensionless (or %)
Xo location of jump of pressure in line crack, m (Ch, 2,9) rJ parameter related to poroelasticity and Poisson ratio, dimensionless (Ch. 3)
Y coordinate, m rJ' viscosity parameter, Pa sn' (Ch. 11)
y dimensionless vertical ordinate for height containment (Ch. 11) computed efficiency from modeler's data, dimensionless (Ch. 8)
Yd location of bottom of perforation, dimensionless (Ch. 11) dimensionless matrix hydraulic diffusivity (Ch. 8)
v; location of top of perforation, dimensionless (Ch, 11) opening time distribution factor, dimensionless
YR vertical coordinate for height containment, m (Ch. 11) Kachanov exponent, dimensionless (Ch. 10)
Yw vertical coordinate for height containment, m (Ch, 11) K compressibility modulus, Pa (Ch. 2)
coordinate, m opening time distribution factor from modeler's data, dimensionless
r foam quality, dimensionless ratio (or %) (Ch. 7) (Ch.8)
Sh; upward height migration, m (Ch. 11) retardation time, s (Ch. 8)
tJ.p pressure drop, Pa interporosity flow coefficient, dimensionless (Ch. 8)
tJ.Pf:ace pressure drop across filter-cake, Pa (Ch. 8) modified interporosity flow coefficient, dimensionless (Ch. 8)
tJ.Ppiz pressure drop across polymer-invaded zone, Pa (Ch. 8) viscosity, Pa . s
tJ.P,e, pressure drop in the reservoir, Pa (Ch. 8) apparent viscosity, Pa- s
6t shut-in time, s (Ch. 9,10) equivalent Newtonian viscosity, Pa s
6t. after-growth time, S (Ch. 9,10) filtrate viscosity, Pa . s (Ch. 8)
6t a, o after-growth time, observed, s (Ch. 9,10) plastic viscosity, Pa s (Ch. 5)
tJ.to dimensionless shut-in time (Ch. 9,10) solvent viscosity, Pa- s (Ch. 5)
6td downward height migration, m (Ch. 11) wall viscosity, Pa- s (Ch. 5)
6p density difference, kg . m-3 (Ch. 7) f.-Lo low shear viscosity, Pa . s (Ch. 5)
e angle of oblique plane, rad (Ch, 2) v Poisson ratio, dimensionless
Cl poroelastic constant, dimensionless auxiliary variable
Cl exponent of fracture length growth, dimensionless (Ch. 9,10) density, kg- rn?
Cl angle characterizing direction of horizontal well, rad (Ch. 3) fluid density, kg m-3 (Ch. 6)
ClKE kinetic energy correction factor, dimensionless (Ch. 5,7) formation density, kg m-3 (Ch. 3)
f3 permeability anisotropy ratio, dimensionless (Ch. 3) normal stress, Pa
f3 auxiliary variable for Carter equation Il, dimensionless (Ch. 9) maximum principal horizontal stress, Pa
f3 angle characterizing direction of horizontal well, rad (Ch. 3) minimum principal horizontal stress, Pa
f3 slip coefficient (Mooney method), mS-1 . Pa-I (Ch.7) maximum stress, Pa
f3c modified slip coefficient (Oldroyd-Jastrzebski method), m2 . 8-1 . Pa-1 minimum stress, Pa
(Ch.7) net stress, Pa (Ch 10)
Y geometry factor relating average to maximum, dimensionless vertical stress, Pa
xx List of Notation

17' effective stress, Pa


17], 172, 0'3 principal stresses, Pa

1
in situ minimum stress in target, upper and lower layer, respectively, Pa
(Ch. 11)
closure quality, Pa (Ch. 3)
r shear stress, Pa
r opening time, s (Ch. 8-11)
dimensionless opening time HYDRAULICALLY INDUCED
yield stress, Pa
volume equalized yield stress, Pa
wall stress, Pa
FRACTURES IN THE
average wall stress, Pa
stress parameter, Pa (Ch. 5)
PETROLEUM AND
porosity, dimensionless
ratio of yield stress to wall stress, dimensionless (Ch. 5) RELATED INDUSTRIES
exponent of width growth, dimensionless (Ch. 8)
fissure porosity, dimensionless (Ch, 8)
matrix porosity, dimensionless (Ch, 8)
angle with the .r axis, rad (Ch. 11)
UJ interporosity constant (Ch, 8)
UJ ratio of inner and outer annulus diameter, dimensionless (Ch. 5)
(I) ratio of particle diameter to half width, dimensionless (Ch. 6)
1.1 Fractures in Well Stimulation
Subterranean porous media have been the source of valuable fluids such as ground-
waters and petroleum, both liquid (oil) and natural gas. Oil and gas, combined, still
account for over 60% of all the energy needs of the world (with coal providing an
additional 30%.) The demand for hydrocarbons is likely to continue unabated at these
exceptionally high levels throughout the twenty-first century (see OPEC data [1]).
Porous formations have also been used for the injection of slurried wastes such as
hazardous chemicals or radioactive byproducts. Certain special geologic structures
have been used for the seasonal storage and quick recovery of already processed
petroleum products and natural gases.
In all of these cases access to the geologic formations has been accomplished
with drilled wells. Historically wells were first vertical with targets of progressively
increasing depth. Then, wells could be drilled deviated and, since the early 1980s they
can be started vertical and after a "buildup" angle they can be turned fully horizontal
into the target formation, with some horizontal lengths exceeding 2500 m (over
8000 ft). Horizontal wells have become commonplace with continuously increasing
estimates on their future share of all wells drilled.
Depths of formations of interest range from a few hundred meters to deeper than
6000 m for natural gas formations. Typical oil reservoirs are usually between 2000
and 3500 m.
Although exact estimations are difficult, it is widely believed that in the USA
alone more than one million petroleum wells have been drilled in the history of the
industry (since 1859 and Col. Drake's well). A comparable number has been drilled
in the former USSR. In the rest of the world the number is smaller.
2 Hydraulically induced fractures Fluid flow through porous media 3

Of the producing wells drilled in North America since the 1950s about 70% of The non-petroleum reader is referred to References [3]-[6] and references therein
gas wells and 50% of oil wells have been hydraulically fractured. The majority of for the developments and solutions to Eqs. 1.3 and 1.5 which are standard in
injection wells have been fractured also (personal communication from Schlumberger petroleum, geothermal and groundwater engineering.
Dowell and Halliburton companies, 1994). Similar percentages are expected in the Of interest are the constant-rate and the constant-pressure-at-the-well solutions.
rest of the world, as those reservoirs mature (age). The general form of the constant-rate solution is
Why is hydraulic fracturing such a common well "stimulation" procedure and
how is it practiced in the modern petroleum and other industries? q = 2rckhD.p (1.6)
These issues are addressed in this chapter and form the rationalization for the fJ-PD
study of hydraulic fracture mechanics. Three different types of flow mechanisms can be distinguished: transient, or
infinite-acting behavior, steady-state with constant outer boundary pressure, Pe, and
pseudosteady-state, denoting a no-flow outer boundary condition.
1.2 Fluid Flow through Porous Media Table 1.1 contains the expressions for the driving pressure gradient D.p and the
dimensionless pressure function, p D, for the three flow mechanisms. Analogous
A porous medium is a geologic formation whose rock contains voids (pores). The expressions can be written for compressible (gas) flow using D.m(p) instead of 6.p
ratio of the pore volume to bulk volume is defined as the porosity, . It is in such a (see Dake [3]; Economides and Ehlig-Economides [6]).
reservoir that fluids are stored. Typical pore diameters range from 10~ 7 m to 10-4 m, Interestingly, for transient rate production at constant Pw r, the solution yields
and reservoir porosities range from about 0.10 to (typical) 0.25 for sandstones to
(extraordinarily high) 0.4 for some carbonate formations. 2:Jrkh(pi - Pwf)
q= 1 ' (1.7)
While the porosity is important in defining the oil- or gas-in-place for a petroleum
fJ--
producing reservoir or the storativity of an injection target, a second quantity, the qD
permeability, k, describing the ability of fluids to flow in the reservoir, is essential. and the PD for constant rate is very nearly equal to the l/qD for constant pressure
The permeability relates the pressure gradient, D.p, which is the driving force in the production for almost all times (see Earlougher [7]).
reservoir with the macroscopic fluid velocity, u, The relationship between q and Pw f and the antecedent engineering activities
u ()( kD.p. (Ll) for their optimum adjustment are the essential functions of petroleum production
engineering (see Economides and Ehlig-Economides [6]).
This is the well known Darcy's law which in radial coordinates yields the following
expression for the volumetric flow rate, q: Table 1.1 Pressure gradients and dimensionless pressure functions for
radial reservoir flow at the well
2:Jrrkhdp
q= ~fJ-~dr' (1.2) D..P PD
Transient Pi - pwf
where fJ- is the viscosity, r is the radial distance, and h is the reservoir thick- (infinite acting reservoir) 1 ( 1)
Po = -2Ez - 4to
ness. Combination of the continuity equation, Darcy'S law and an equation of state,
describing incompressible fluid, yields the well known diffusivity equation kt
and to= ~-,-2-
4>I-LC, w

[pp 1ap fJ-Cr ap Semilogarithmic Po = ~(IntD+0.8091)


-+--=---, (1.3)
or2 r ar k at approximation at to > 100

where c, is the total system compressibility and t is the time. Steady state P. - pwf PD= In!j_
rw
An analogous expression for gas (compressible) flow employs the real-gas pseu- O.472re
Pseudosteady state p- Pwf po=[n--
dopressure, m(p), defined by Al-Hussainy and Ramey [2] as rw
p 2p Ei = Exponential integral
m(p) =
lPO
-dp,
fJ-Z
(1.4) Pi
Pe
= Initial reservoir pressure
= Outer boundary constant pressure
p = Average reservoir pressure
and, thus, pwf = Flowing bottom-hole pressure
a2m(p) 1 am(p) fJ-Cr am(p) = Outer boundary radius
-- + --- = -----. (1.5) Ye

or2 r ar k at rw = Well radius.


.... _-_ ----------------------- -----_ ..........-._-----_. __ ..._-----

4 Hydraulically induced fractures Flow from a fractured well 5

1.2.1 The Near-well Zone high 9.87 x 10-14 m1 (100 md) and Pwf =2 X 107 Pa. Use the steady-stateexpression
of Eq. i.s.
Converging radial flow de facto exaggerates the impact of the near-well zone. It is
clear from Eq. 1.6 that for (e.g. steady-state) flow, the driving pressure gradient is
proportional to the logarithm of the radial distance.
Solution
An alternative way to state this is that for a constant production rate, the same For k = 9.87 X 10-14 m2, Eq, 1.8 yields
amount of pressure gradient is consumed in the first meter as in the next 10 m,
the next 100 m, etc. Thus, by analogy, it should be obvious that alterations to the (2)Jr(9.87 x 10-14)(20)(3.5 x 107 - 2 X 107) 0.186
natural permeability in the near-well zone would be critical to the well production q= [300] =8+s'
(1 x 10-3) In- +s
or injection rate at constant tlp. 0.1
Permeability-altering phenomena occur frequently in almost all well operations
and for s = 10, q = 1.03 X 10-2 m3/s whereas for s = 0, q = 2.32 X 10-2 m3/s.
including drilling, well completions or "workovers", Reduction of the near-well reser-
Both the incremental flow rate (1.29 x 10-2 m3/s = 7010 barrels/day) and the
voir permeability is common, is referred to as damage, and has been characterized post-treatmentrate itself (2.32 x 10-2 mJ /s = 12600 barrels/day) are very attractive,
by a dimensionless skin effect, s (see Van Everdingen and Hurst [8]) analogous to pointing towards matrix stimulation.
the film coefficient in heat transfer. Assuming that a minimum well production rate equal to 9.2 x 10-5 m3/s
This skin effect, implying a steady-state pressure drop, is added to the dimension- (50 barrels/day) is required, then from Eq. L8, with s = 0, the minimum reservoir
less pressure in Eq. 1.6, resulting in a change in the well production or injection rate: permeability for which matrix stimulation is attractive would be k = 3.9 X 10-16 m2
(0.4 rnd), In production engineeringthe attractiveness of the stimulation is subject to
2rrkhtlp the costs of the treatmentwhich must be balancedagainst the benefitsof the incremental
(1.8)
q = J.i(PD +s)" production rate of 5.1 x 10-5 m3/s (28 barrels/day).
In this exercise, for perrneabilitiesless than 3.9 x 10-16 m2 (or in some cases for
The reader is referred to Chapter 5 of Economides et al. [6] for an extensive much higher permeabilitiesif economic considerationsindicate) hydraulicfracturing is
description of the various causes of near-well damage, certain mechanical contribu- likely to be the appropriatewell stimulationoperation. 0
tions to the skin effect and quantification of its impact.
The skin effect is determined through the pressure transient testing of a well. A
large and positive value implies damage or a flow impediment due to a mechanical 1.3 Flow from a Fractured Well
reason (e.g. s = 20), whereas s = 0 is for undisturbed permeability in a vertical welL
Once a hydraulic fracture is created in a well or, in the not uncommon case, where
Zero skin could imply damage in a deviated well. A negative skin implies stimu-
the well intersects a natural fracture, fluid will flow normal to the fracture face from
lation where the near-well permeability is larger than the original reservoir value.
or to the reservoir (production or injection) and then along the fracture path from or
The latter case can be accomplished through matrix stimulation, which includes a
to the welL
series of possible chemical treatments intended to remove near-well damage once
For almost all depths of interest (as will be expounded upon in detail in Chapter 2)
its nature is identified (see Economides et al. [6]). Larger post-stimulation perme-
a hydraulic fracture will be largely vertical. Gringarten and Ramey [9] have described
abilities are possible, although rare. This could happen if the formation itself reacts
the flow performance of an infinite-conductivity fracture whereas Cinco-Ley and
with the injected stimulation fluids (e.g. a hydrochloric acid, HCI, solution and a
Samaniego [10] dealt with the finite-conductivity fracture case. The latter is a reason-
carbonate rock).
able description of created hydraulic fractures.
Hydraulic fracturing may be attempted in those cases where matrix stimulation
In the case of an infinite-conductivity fracture (the upper limit of high conduc-
cannot result in an economically satisfactory well production or injection rate.
tivity) flow of fluid is characteristically linear, i.e., from the reservoir into the fracture.
To understand the need for an alternative to matrix stimulation the following
Once the fluid enters the fracture, it is presumed to enter the wellbore instantaneously,
example is offered.
relative to the time it would take without the fracture.
For the finite-conductivity fracture a discernible linear flow develops within the
Example 1.1 Matrix Stimulation vs. Hydraulic Fracturing fracture, in addition to the linear component from the reservoir into the fracture,
Suppose that a well with 'w = 0.1 m is drilled in a reservoir with r = 300 m, h = hence the characteristic term bilinear flow (see Cinco-Ley and Samaniego [10]).
20 m and Pe = 3.5 X 107 Pa. If the fluid viscosity f.L = 1 X 10-3 Pa s and a well test Figure 1.1 is the Cinco-Ley and Samaniego [10] solution, as plotted by Agarwal
has provided s = 10, investigate the incremental well flow rate before and after a et al. [11] for the transient flow of a finite-conductivity fractured well. On the
completely successful matrix stimulation(i.e. with s = 10 and s = 0, respectively)for ordinate is the dimensionless pressure, PD, on the abscissa is the dimensionless
a range of permeabilities from a low value equal to 9.87 x 10-18 m2 (0.01 md) to a time, tD:cf and the parameter is the dimensionless fracture conductivity, F CD.
----~-- ..-.-.....__ ---

Hydraulic fracture design 7

2.5
\
\
~....
2
"::..
C
E.
+
",-
1.5

1\
I"
t-

0.5
10' 100 10' 10' 10'

Dimensionless Time, tDXf


Figure 1.1 Finite-conductivity fracture solution. Dimensionless pressure vs. dimensionless Figure 1.2 Equivalent skin effect for pseudoradial flow into a fractured well [101
time [111

These are defined for liquid (oil) as: typical fracture permeability, kl, is 9.87 x 10-11 m2 (100000 md) and the propped
2rrkh(Pi - Pwf) fracture width is 5 x 10-3 m. Calculate the steady-state production rate if the fracture
PD= , (1.9) half-length is 300 m.
qj.
kt
tDx! = ---2' (1.10)
MCrXf Solution
kfw
and FCD = kxr ' (1.11) From Eq. 1.11 F CD =
4.2 and, therefore, from Figure 1.2, sf + In(xI /rw) = 0.96.
Substituting the values of XI and rw (= 0.1 m) this would lead to 51 = -7.
In Eqs. 1.9 to 1.11 variables are as defined in Eq. 1.6 and Table 1.1, except for Using Eq. 1.8, for steady-state production and S I = =
-7 results in q 7.35 X
10-4 m3/s (400 barrels/day) which is an 8-fold increase over the best case that this
the fracture half-length, x f' the fracture permeability, k f' and the propped fracture
well would produce with matrix stimulation (i.e. s = 0).
width, w.
It is essential to note that once a well is hydraulically fractured tbe overwhelming
The values of the fracture half-length and fracture conductivity are the essential portion of the total flow is through the fracture, bypassing the damage zone and, thus,
quantities for the prediction of fractured well performance. any pretreatment radial skin effect can be ignored. 0
The Cinco-Ley and Samaniego [10] solution becomes indistinguishable from the
Gringarten and Ramey [9] solution for F CD > 300. For practical purposes they can
be considered as the same for F CD > 70. 1.4 Hydraulic Fracture Design
Long-term fractured well performance results in pseudoradial flow, and the
presence of a hydraulic fracture of half-length, x f and conductivity, F CD, can be The proper engineering approach to hydraulic fracture design is to maximize the
manifested by an equivalent skin effect, sf, which can be read from Figure 1.2. post-treatment performance and ensuing benefits at the lowest treatment costs. Thus,
an economic criterion such as the net present value (NPV) has been employed for
Example 1.2 Performance of a Fractured vs. an Unfractured Well this purpose: the optimum fracture size would coincide with the maximum NPV (see
Meng and Brown [12]).
Suppose that the well in Example 1.1 with permeability k = 3.9 X 1O-l6 m-, and for
A common hydraulic fracture design optimization procedure starts from a fracture
which matrix stimulation has been deemed unattractive, is hydraulically fractured. A
size, usually denoted by, but not limited to, the fracture half-length.
8 Hydraulically induced fractures Hydraulic fracture design 9

A fracture-propagation model then describes the hydraulic fracture geometry defi- fracture.) Use a realistic fracture permeability, taking into account possible damage to
nitely including the width and, with an appropriate model, the fracture height. This the pr~ppant~kj = 1 X 10-11 m2. Assume that the created fracture height equals the
issue is addressed in detail in Chapters 9-11. formation thickness. Use the Cinco and Samaniegograph, Figure 1.2, which assumes
pseudoradialflow.
The required fracturing fluid volume is then estimated through a material balance
accounting for the created fracture volume and the fluid leakoff normal to the fracture
faces. This calculation simultaneously provides the required injection time. Solution
Chapter 8 contains fracture leakoff models and the manner in which they are
incorporated in the fracture-propagation material balance. The.same propped volume can be establishedcreating a narrow, elongated, fracture or
There are several techniques to estimate the required mass of proppant materials. a Wide but .short one. The production rate will depend on the decision according to
The calculation depends on the manner of propp ant addition to the fracturing fluid Eq. 1.8, which for steady-stateproduction rate takes the form
slurry. A common method suggests a rampedproppant schedule (see Nolte (13]) with
its onset depending on the leakoff characteristics. Thus, after the end of injection
the mass of proppant leads to the propped fracture width assuming that the fracture
length is either equal to the hydraulic length or it is truncated by some practical
criterion, e.g. where the width is equal to three proppant diameters. The choice
of proppant is critical since the fracture permeability at the expected in situ stress Th~pse~dor~dial,steady-s~a~eflowim~liedby this relationshipshould emergerelatively
depends on the strength of the proppant (see Brown and Economides [14]). rapidly III higher-permeability formations, which are the normal candidates for "frac
Thus, the propped width, w, the fracture permeability, k I> the assumed fracture & pack" treatments. Obviously, our aim is to minimize the denominator. This can be
half-length, XI and the reservoir permeability, k are sufficient to allow the forecast of accomplished using the Cinco and Samaniego graph, which is a plot of the function
f), defined by
the post-treatment well performance using the model presented in Section 1.3. This
prediction leads readily to the future incremental benefits which, when discounted f I(Iogro F CD) = Sf + In Xf.
to the present, constitute the net present value of the incremental revenue. r",

Inherent to this design procedure is the estimation of the required fluid volume, We will use th~function f 1. replotted in Figure 1.3 for convenience.Given the function
proppant mass and time of injection. These are the main components of the treatment f I the denominator of the production rate can be expressed as
costs which, when subtracted from the present value of the incremental revenue, lead
to the NPV, specific for the assumed fracture half-length. r. xf
In - -In -
Tw rw
+ fl(lOglo F CD),
The procedure is then repeated with increments of the fracture half-length and for
each the corresponding NPV is determined. Optimum xI is the one corresponding
5
to the maximum NPV.
In an appropriate engineering design it is this treatment that should be executed.
Typically indicated half-lengths may range from less than 100 m for a higher perme-
4
!
ability reservoir to more than 500 m for a low-permeability formation. 3 v -1
With the advent of the tip screen-out technique ("frac & pack"), especially in high-
permeability, soft, formations, it is possible to create short fractures with unusually
~
2 "....
I'.. .,. / ~
wide propped width. In this context a strictly technical optimization problem can be ~ I'- V~ J
formulated: how to select the length and width if the propped fracture volume is -e: '1
given. Example 1.3 deals with this problem. 0
/1--'

,...
1
Example 1.3 Optimal Fracture Conductivity
2
Consider once again the reservoir and well data of Example 1.2 (k = 3.9 x
10-'
10-16 m2, h = 20 m, r. = 300 m, J.l = 1 X 10-3 Pa- s, Pe = 3.5 X 107 Pa and 10 10' 10" 10'
pwf = 2 x 107 Pa). Determine the optimum fracture half-length, Xj, the optimum FCD
propped width, w, and the optimum steady-state production rate if the volume of the
Figure 1.3
propped fracture, V f = 100 m3, is given. (Note that V f is the volume of the two-wing Functionsfor optimalfractureconductivityas usedin Example 1.3
----------- ------
.. ----_._ ._---- ..._---_.

10 Hydraulically induced fractures Treatment execution 11

which can be further simplified to give The optimum production rate (assuming P. = 3.5 X 107 Pa and pwl =2 x
107 Pa) is

16 20(3.5 X 107 - 2 X 107)


2]I' X 3 ,x
9 10 - X -'-------,:-;:---'-
From the above expression we can eliminate the half-length using the relationship, 1 X 10-3
q = ~---r======~==---
Vf = Zwhx]; and the definition of the fracture conductivity, Eq. 1.11. As a result, we 100 x 10-11
arrive at the following minimization problem:
In 300 -In =---=-=-=-=--:-=-~
2 x 20 x 3.9 X 10-10
+ 1.45
= 4.54 X 10-3 m3/s(247 barrels/day).

In general it is necessary to check if the resulting half-length is less than r, (otherwise


xf has to be selected to be equal to r.). Similarly, one has to check if the resulting
where the only unknown variable is F CD. The first two terms are constant, and hence do
optimum width is realistic, i.e, it is greater than, say, three times the proppant diameter
not affect the location of the minimum. The last two terms do not contain any problem-
(otherwise a threshold value has to be selected as the optimum width.) In our example
specific data. Therefore, the optimum F CD is a given constant for any reservoir, well
both conditions are satisfied.
and proppant. (Moreover, the same optimum F CD would result for pseudo steady state
The above example provides a deeper insight into the real meaning of dimensionless
production rate.)
fracture conductivity. The reservoir and the fracture can be considered as a system
To find the optimum F CD we introduce two new functions: the first one, denoted by
working in series. The reservoir can deliver more hydrocarbon if the fracture is longer
12, is defined by but with a narrow fracture the resistance to flow may be Significant inside the fracture
itself. The optimum dimensionless fracture conductivity (F CD.opr 1.2) corresponds to=
the best compromise between the requirements of the two subsystems. 0
and is plotted as a straight line in Figure 1.3.
The function f 3, which we wish to minimize, is simply the sum of f 1 and f 2 As
seen from Figure 1.3 it has a minimum at F CD.opr =
1.2 where f 3,opr =
1,45. Therefore, 1.5 Treatment Execution
the following results hold:
The optimum half-length is given by Hydraulic fracturing is a massive operation, frequently resulting in the injection
of more than 2000 m3 of fracturing fluids, 5 x lOS kg of proppants at bottomhole
rv;k; pressures that could be over 5 x 107 Pa (corresponding to wellhead pressures of
xf = V 2.4hk' 2 x 107 Pa) while employing as many as two dozen active or standby pumping units
each capable of delivering 1500 to 2000 hhp (1100 to 1500 kW). Analogous power
the optimum width is obtained from may be available on specially designed stimulation vessels for offshore operations.
Figure 1.4 is a schematic depiction of the execution operation. Fracturing fluids
with appropriate additives are blended with metered proppant and then injected
through appropriate fracturing strings into the target formation. Below, there is
a brief description of fracturing fluids, their expected performance, the additives
and the optimal steady state production rate is that affect this performance and common propping materials. Brown and Econo-
mides [14] contains a much more detailed description along with large amounts of
21ikh~p
data required for the selection of fluids and proppants. Chapters 5 to 7 of this book
_ f.L
q- [V;k; describe the rheology and fluid mechanics of fracturing slurries.
In r. - In V -!if + 1.45
1.5.1 Fracturing Fluids
Returning to our numerical example the following results are readily calculated:
Fracturing fluid properties are expected to facilitate fracture initiation (breakdown),
XI = 100 X 10-11 = 232 m, fracture propagation and proppant transport while they minimize leakoff and long-
2.4 x 20 x 3.9 X 10-16 term residual damage to the proppant-pack permeability.
Viscosity is, thus, the essential property and may be augmented by additives during
w=
0.6 x 100 x 3.9 X 10-16 = 0.011 m.
20 x 10-11
execution. It must be destroyed by other additives after the treatment.
12 Hydraulically induced fractures Treatment execution 13

have considerably reduced viscosities (e.g. < 2 x 10-2 Pa- s) which are insufficient
for proppant transport. Required minimum viscosity in the fracture, where large shear
rates at the tip may reduce the viscosity further, is considered to be 0.1 Pa- s (see
Brown and Economides [14]).
To increase the viscosity substantially, crosslinkers of the polymer chains have
been employed. For temperatures below 115C borate crosslinkers are considered
desirable. For higher temperatures, organometallic crosslinkers such as titanium and
zirconium complexes are necessary. To meet the demand for lower viscosity in the
tubulars and higher viscosity in the fracture, delayed crosslinkers have been used.
These are triggered by activators that are sensitive to the high-shear values as the
fluid passes through the perforations. To avoid oxidative degradation in the fracture,
oxygen scavengers are often added to the fluid.
A "40-1b borate-crosslinked gel" (40 lb/Mgal = 4.8 kg/nr') at a reservoir temper-
ature of 90C would still have an apparent viscosity of 0.2 Pa . s after 4 hours of
injection-induced shear (see Brown and Economides [14]).
Oil-based fluids have been used in water-sensitive formations with a phosphate
ester as the gelling agent. These fluids are losing their "market share" because of
environmental and obvious safety considerations.
Focus of research has been the development of non-intrusive, non-damaging water-
based fluids.
A very common practice is the foaming of fracturing fluids with carbon dioxide
or nitrogen. Foam qualities (gas volume fraction) from SO to 90% have been used
with 70 being very common. The purpose in using these fluids is to minimize filtrate
damage and, more importantly, to facilitate the cleanup: fluid fiowback after the
treatment.
After the injection stops the formidable task of breaking down the polymer
emerges. Unbroken polymer chains result in a marked reduction in the permeability of
Figure 1.4 The fracturing operation. Fracturing fluids and proppants are blended and injected . the proppant pack. Thus, oxidizers or enzymes, and at times encapsulated breakers,
downhole at the target formation
are added to the fracturing fluid. The breaking action is critical to the success of
hydraulic fracturing and is the subject of active ongoing research.
The ideal fluid has low viscosity in the horizontal and vertical tubulars to reduce
the friction pressure and, therefore, the required treating pressure. After the fluid
enters the fracture, the viscosity should have a high value to cause a larger width and 1.5.2 Proppants
better proppant transport. In addition, the same agents that enhance viscosity may be
used for the building of a filtercake on the fracture walls to reduce leakoff. After the The hydraulic width created during the injection is reduced to zero after supplied
treatment, the high viscosity is no longer needed but, instead, it is highly detrimental fracturing pressure subsides to the closure pressure, unless propping materials are
to the flow of produced or injected fluids. Thus, it must be reduced considerably. used. It is this residual propped width that can be used for the forecast of fractured
These contradictory functions are essential elements in the fracturing fluid design. well performance that was outlined in Section 1.3.
Fracturing fluids have been based on water, oil, mixed water and oil (emulsions), Proppant size and proppant strength are the main criteria for selection. The general
mixed water and gas or mixed oil and gas (foams). families of propp ants are divided into low, intermediate and high strength. The
For water-based fluids, common polymer thickeners are hydroxyethyl cellulose demand for strength is directly related to the level of stress that the proppant will
(REC) and hydroxypropyl guar (HPG) in quantities varying (in field units) from experience in the long term.
20 lb/Mgal (2.4 kg/nr') to 80 lb/Mgal (9.6 kg/nr'). At ambient conditions these Low-strength proppants are natural sands in typical sizes from 12/20 mesh to
polymer solutions may lead to viscosities up to 0.1 Pa- s (at expected shear rates 20/40 mesh (average particle diameter is 2 x 10-4 m to 1 x 10-4 m). They are
in a fracture) but at reservoir temperatures (T = 6SC to l1SoC or even higher) they usually attractive at depths less than 2000 m because although they are the least
------._ ...----.-- .. ---.- -----_ .... . --------_ .. _-----------------

14 Hydraulically induced fractures


Mechanics in hydraulic fracturing 15

expensive propp ants they undergo severe crushing resulting in substantial proppant- wellbore deformations, corresponding to stress anisotropy before the treatment and,
pack permeability reduction. An NPV-based design procedure allows the balancing potentially, the stress induced after a treatment.
of these effects and is an invaluable aid in deciding on the appropriate proppant.
Frequently, sands are coated with resins which allow the fragments to stay together
and thus maintain a high fracture permeability at larger stress values. 1.6.2 Core Measurements
Synthetic, intermediate- and high-strength proppants are used at depths up to 3000
and 5000 m, respectively. The appearance and disappearance of fissures as the stress on a core is reduced or
Brown and Economides [14] contains an extensive coverage of proppant proper- increased and the counting of these fissures has been used for the determination of
ties including their degradation from long-term exposure to stresses. stress anisotropy in oriented cores.
Strain relaxation and its measurement with sensitive devices has been referred to
as anelastic strain recovery (see Blanton [25]; Teufel [26]).
1.6 DataAcquisition and Evaluationfor Oriented cores are specially prepared and fitted with gauges which detect the
Hydraulic Fracturing relative displacement resulting from strain recovery.
The reverse procedure is used for the differential strain recovery analysis where
Field and laboratory measurements are often conducted before and after a hydraulic cores are re-stressed and the relative differences in displacements are correlated with
fracture treatment to predict and evaluate fracture geometry and conductivity. stress anisotropy (see Strickland and Ren [27]).
Data acquisition involves well logging, core laboratory investigations, well testing
and fracture calibration injections. Seismic techniques, although expensive, can be
1.6.3 Well Testing
used in critical cases. The data acquisition has a cost and, thus, the selection of
tests depends on the benefits from the knowledge of particular variables and the Pressure transient testing is widely practiced by engineers dealing with porous media.
opportunity cost of their ignorance. Analysis of the pressure and rate data while the well is flowing (drawdown) or shut
Appropriate selection of data acquisition techniques is an essential part in the in (buildup) or observed by another well (interference) allows the determination
success of hydraulic fracture design. of important well and reservoir variables. These include the skin effect, reservoir
Although this book falls outside the scope of data acquisition and evaluation, permeability and permeability anisotropy, types and locations of boundaries and
below is an account of common techniques complete with appropriate references for formation heterogeneities (such as two-porosity systems.)
further reading. For wells that could be candidates for hydraulic fracturing, a pretreatment well
test can reveal the reservoir permeability and skin effect allowing a decision for
stimulation (matrix vs. fracturing vs no treatment at all.) If fracturing is indicated the
1.6.1 WellLog Measurements
reservoir permeability is a critical variable for the design optimization (see Balen
Pretreatment log measurements are intended to obtain mechanical properties of et at. [28]).
the target and adjoining intervals and predict stress values and, especially, stress A post-treatment well test, and assuming the reservoir permeability is known, can
contrast. This would give indications for the fracture height migration (see Newberry provide the fracture half-length and fracture conductivity. Such a determination is
et al. [15]; Ahmed et al. [16]). essential for the design evaluation.
Borehole acoustic televiewers are used for the measurement of sonic travel time In Chapter 11 of Reference 6, modem well test analysis techniques are presented,
and amplitude (see Pasternak and Goodwill [17]; Plumb and Luthi [18]). Identifying complete with well-test design guidelines and types of tests that are presently prac-
borehole ellipticity and the presence of vugs and natural fractures provide evidence ticed in the industry.
of stress anisotropy and, thus, the expected hydraulic fracture azimuth.
Mechanically fitted dipmeter logs with four and six arms are used to detect
open-hole ellipticity and stress-related wellbore breakouts. These effects have been
1.7 Mechanicsin Hydraulic Fracturing
correlated clearly with stress anisotropy [19-21]. Rock, fracture and fluid mechanics are critical elements in the understanding and
Dipmeter logs with a dense array of microresistivity detectors provide wellbore engineering design of hydraulic fracture treatments.
images where natural fissures can be mapped. These devices are used in both vertical Rock mechanical properties dictate the stress and stress distribution at depth
and horizontal wells [18,22,23]. (Chapter 3) and elastic properties control the created fracture geometry (Chapters 2
More recently, a downhole extensiometer has been introduced by Lin and Ray [24] and 4). Contrast between the properties of adjoining layers controls the vertical
with two six-arm calipers and very sensitive pressure transducers to detect small fracture height migration (Chapter 11).
-----, .'_-,_._------_-_ ----

16 Hydraulically induced fractures References 17

Fracture mechanics is an obvious field of study in this endeavor allowing for the 17. Pasternak, E.S. and Goodwill, G.D.: Application of Digital Borehole Televiewer
interaction between the provided pressure and the resisting stresses. Tip propagation Logging, Proc. 24th Annual SPWLA, 1983.
mechanisms and their effect on the observed net pressures are subjects of ongoing 18. Plumb, R.A and Luthi, S.M.: Application of Borehole Images to Geologic Modeling of
an Eolian Reservoir, Paper SPE 15487, 1986.
research and controversies (Chapters 10 and 11).
19. Brown, R.O., Forgotson,l.M. and Forgotson, I.M., Jr.: Predicting the Orientation of
The combination of rock, fracture and fluid mechanics results in the study of frac- Hydraulically Created Fractures in the Cotton Valley Formation of East Texas, Paper
ture propagation, the interaction and sensitivity between treatment variables and the SPE 9269, 1980.
formation to be fractured and the resulting hydraulic fracture morphology. These 20. Gough, D.I. and Bell, 1.S.: Stress Orientations from Oil-Well Fractures in Alberta and
concepts are treated extensively in Chapters 9 to 11. They are also the central Texas, Can. Jour. Earth Sci., 18, 638-645, 1981.
elements of this book. 21. Zoback, M.D. and Zoback, M.L.: in Neotectonics, G.S.A, 1988.
22. Svor, T.R. and Meehan, D.N.: Quantifying Horizontal Well Logs in Naturally Fractured
Reservoirs - I, Paper SPE 22634, 1991.
References 23. Meehan, D.N. and Svor, T.R.: Quantifying Horizontal Well Logs in Naturally Fractured
Reservoirs - II, Paper SPE 22792, 1991.
24. Lin, P. and Ray, T.G.: A New Method to Determine In-Situ Stress Directions and In-Situ
1. Anonymous, OPEC's Facts and Figures, Organization of Petroleum. Exporting
Countries, Vienna, 1993. Formation Rock Properties During a Microfrac Test, Paper SPE 26600, 1993.
25. Blanton, T.L.: The Relation Between Recovery Deformation and In-Situ Stress Magni-
2. Al-Hussainy, R. and Ramey, H.I., Ir.: Applications of Real Gas Theory to Well Testing
tudes, Paper SPE 11624, 1983.
and Deliverability Forecasting, JPT, (May), 637-642, 1966.
26. Teufel, L.W.: Prediction of Hydraulic Fracture Azimuth from Anelastic Strain Recovery
3. Dake, L.P., Fundamentals of Reservoir Engineering, Elsevier, Amsterdam, 1978.
4. Craft, B.C. and Hawkins, M. (Revised by Terry, R.E.) Applied Petroleum Reservoir Measurements of Oriented Cores, Proc. 23rd U.S. National Rock Mechanics Symposium
1982. '
Engineering, 2nd ed., Prentice Hall, Englewood Cliffs, NJ, 1991.
27. Strickland, F. and Ren, N.: Predicting the In-Situ Stress of Deep Wells Using the Differ-
5. Amyx, J.W., Bass, D.M. and Whiting, R.L.: Petroleum Reservoir Engineering; Physical
ential Strain Curve Analysis, Paper SPE 8954, 1980.
Properties, McGraw Hill, New York, 1960
28. Balen, R.M., Meng, H.-Z. and Economides, M.J.: Application of the Net Present Value
6. Economides, M.J., Hill, AD. and Ehlig-Bconomides, C.A: Petroleum Production
(NPV) in the Optimization of Hydraulic Fractures, Paper SPE 18541, 1988.
Systems, Prentice Hall, Englewood Cliffs, N.J., 1994.
7. Earlougher, R.C., Jr.: Advances in Well Test Analysis, SPE, Dallas, TX, 1977.
8. Van Everdingen, AF. and HUrst, N.: The Application of the Laplace Transformation to
Flow Problems in Reservoirs, Trans. AlME, 186305- 324, 1949.
9. Gringarten, AC. and Ramey, A.J., IT.: Unsteady State Pressure Distributions Created
by a Well with a Single-Infinite Conductivity Vertical Fracture, SPEl, (Aug.), 347-360,
1974.
10. Cinco-Ley, H. and Samaniego, F.: Transient Pressure Analysis for Fractured Wells, JPT,
1749-1766,1981.
11. Agarwal, R.G., Carter, R.D. and Pollock, C.B.: Evaluation and Prediction of
Performance of Low-Permeability Gas', Wells Stimulated by Massive Hydraulic
Fracturing, JPT (March), 362-372, 1979; Trans. AlME, 267.
12. Meng, H.Z. and Brown, K.E.: Coupling of Production Forecasting, Fracture Geometry
Requirements and Treatment Scheduling in the Optimum Hydraulic Fracture Design,
SPE Paper 16435, 1987.
13. Nolte, K.G.: Determination of Proppant and Fluid Schedules from Fracturing Pressure
Decline, SPEPE, pp. 255-265, July 1986.
14. Brown, J.E. and Economides, M.J.: Practical Considerations in Fracture Treatment
Design, in Economides, MJ.: Practical Companion to Reservoir Stimulation, Elsevier,
Amsterdam, 1992.
15. Newberry, B.M., Nelson, R.F. and Ahmed, U.: Prediction of Vertical Hydraulic Fracture
Migration Using Compressibility and Shear Wave Slowness, Paper SPE/DOE 13895,
1985.
16. Ahmed, U., Newberry, B.M. and Cannon, AM.: Fracture Pressure Gradients
Determination from Well Logs, Paper SPE/DOE 13857, 1985.
-------_-----_------._------ __ -- ----- ------ -----------._----

2
LINEAR ELASTICITY,
FRACTURE SHAPES
AND INDUCED STRESSES

A purely elastic body has a natural state to which the body returns if all the external
forces are removed. An elastic deformation is therefore reversible: The work done
on the body is saved as elastic energy which is totally recoverable. If deformations
and their inducing forces (or forces and their inducing deformations) are connected
by a linear relationship, this is linear elasticity. The appearance and propagation of
a fracture means that the material has responded in an inherently non-elastic way
and an irreversible change has occurred. At first glance, therefore, it seems that
elastic theory (linear or even non-linear) might be of little use in fracture mechanics.
Nevertheless, linear elasticity is a useful tool when studying fractures, because both
the stresses and deformations (except for the fracture surface and perhaps the vicinity
of the tip) may be still well described by elastic theory.

2.1 Force and Deformation


Forces considered in elastic theory (see Billington and Tate [1]; Fenner [2]) are
distributed by nature. Surface forces are distributed along a surface and body forces
along a volume. In both cases what really matters is the intensity, i.e. the force acting
on a unit area of the surface or in a unit volume of the material. The action of the
surrounding material on any volume element of it is transmitted by surface forces
and thus, we concentrate on them.

2.1.1 Stress
The ratio of the force to the elementary surface area it is acting on is the force
intensity called stress (or surface traction):

a = lim
IlA .... O
(D.F)
M
, (2.1)

measured in N/m2 or, briefly, Pa,


20 Linear elasticity Force and deformation 21

The stress is a vector with magnitude and direction. An elementary 'surface is are CT;m CTyy, Uu., .xy, ryz and .zx. The remaining three components are given by the
contained in a plane which can be rotated arbitrarily and hence there is an infinite restrictions:
set of stress vectors associated with a given point The stress state is given if we (2.2)
provide an appropriate means to determine the stress corresponding to any arbitrarily
If the six independent stresses are specified, the stress acting on any arbitrarily
selected plane direction.
oriented (oblique) plane can be obtained by applying force balance. The word
Stresses normal to the plane may be tensile or compressive, while those parallel
"obtained" means that we can calculate the three components of the stress vector.
to the plane are called shear. A normal stress is readily visualized based on everyday
(The actual expressions will be given later.) Once the stress vector is known, we can
experience. To understand shear stress properly some abstraction is needed. Any
decompose it into a normal and a shear component relative to the specified plane.
stress can be decomposed into two orthogonal shear components and a tensile (or Given the state of stress at a point, we may continuously change the orientation of the
compressive) one. A common system of notation includes two suffixes: the first one oblique plane while the magnitudes of the normal and shear .stresses are v~rying. It
refers to the direction of the stress while the second one denotes the direction of the happens that there are three specific orientations where the shear stress vanishes and
outward normal to the plane on which it acts. A tensile stress (positive by convention) (at the same time) the normal stress has a local maximum. The three local maxima
and a compressive stress (negative) have two identical suffixes. Shear stresses have are called principal stresses. The three eigenvalues of the matrix
different suffixes. To emphasize the difference, shear stresses are often denoted by
r. If there is no danger of misinterpretation, the second suffix of a normal stress can
be deleted (since it is identical to the first one.) (2.3)
Figure 2.1 shows an elementary cube whose edges are parallel to the Cartesian
coordinate axes. There are nine stress components but they cannot be selected inde- denoted by CT1 :::: CT2 :::: 0'3 give the magnitude of the principal stresses. The compo-
pendently. Rotational equilibrium poses three constraints on them. The state of stress nents of the corresponding eigenvectors are the direction cosines of the plane (with
at a point is determined by six independent stresses: In Cartesian coordinates these respect to which the maximum occurs) and hence components of the direction vector
of the principal stresses. Moreover, the eigenvectors are mutually orthogonal (a
consequence of the symmetry of the matrix).
In some applications the eigenvectors provide a natural coordinate system. In this
coordinate system the matrix (2.3) will be diagonal. The eigenvalues of a diagonal
matrix are the diagonal elements. If we know the directions of the principal stresses,
then the only three additional data needed to specify the stress state are the diagonal
elements of the matrix, i.e, O'h CT2 and 173.
In geologic applications often we may assume that one of the principal stresses
is vertical. Then one additional angle has to be given to specify the direction of the
second principal stress in the horizontal plane. The third principal direction is also
horizontal and orthogonal to the second one. Since such a direction is, unique the
only additional data we need are the values 0'1, 0'2 and 0'3

2.1.2 Strain
We can think of a deformation as a transition from a reference configuration into
another one. Simple translation or rotation of a rigid body are also deformations, but
are of little interest in the present context. In elasticity theory the interest is with
deformations, where the relative position of the points changes.
For defining a suitable measure of the deformation let us consider two material
points. If I is the original distance between the two points and I + Sl is the new
distance, the engineering strain is defined by
t:.l
e=-. (2.4)
Figure 2.1 Stresses acting on one surface of an elementary cube I
._------------_---

22 Unear elasticity Material properties 23

Tensile strain corresponds to extension whereas compressive strain corresponds to Hence, a suitable definition of the first component of the strain state, ex" in accor-
contraction. By convention, strain associated with extension is negative and compres- dance with Eq. 2.4 is
sive strain is positive. However, in rock mechanics and especially in hydraulic
fracturing sometimes the opposite convention is more appropriate. The actual sign (2.7)
convention should be clear from the context. A shear strain is associated with plane
layers sliding over each other. For small strains the angle of distortion (in radians) Similar arguments lead to the definition of other strain components listed in Table 2.1.
is a suitable measure of the shear strain. Again, six independent components (en, eyy, eZZ'exy, e}Z and ezx) should be spec-
For the full definition of strain in the three-dimensional space, it is necessary ified to give the state of strain at a given point.
to consider a point in the original configuration with coordinates x, y and z. After
deformation, the new coordinates will be x + ux, y + uy and z + u-, respectively (see
Figure 2.2). The quantities u, uy and Uz are the components of the displacement 2.2 Material Properties
vector. With changing location of the original point the displacement may vary but
smoothly. If we consider a straight line starting from (x, y, z), parallel to the x_ axis Real materials have complex behavior when subjected to a stress field. Idealized
and of length ox (where this length is short enough) then models help to understand the main features of the behavior. A perfectly elastic
material stores the work done on it by external forces, and then it allows full recovery.
I = ox, (2.5) How an elastic material responds with strain to a specific stress state (or vice versa)
can be described by a constitutive equation. Of particular interest is the case where
and the constitutive equation is linear.
1+tll
aux
= ox+ -ox. (2.6)
ax
2.2.1 Linear Elastic Material
For a linear elastic material the stress varies linearly with strain. Hooke's law states
(X' + tt.. y' + U~)
that under uniaxial compression the stress induced is proportional to the strain

(2.8)

where E is Young's modulus. As shown in Figure 2.3, the deformation in the x


(x', y) direction is accompanied by an additional deformation in the y direction. This "side

Figure 2.2 Displacement and strain

Table 2.1 Strain components, Cij

Secondindex x y z
First index

x
aux
ax
1
2
(aux +auy)
oy ax -
I
2
C
u: au,)
-+-
ax oz
y see xy au}'
ay -
I
2
Cazu}'
-+-au,)
oy e
v=-cyy
yz au, xx
z see .rz see
8z Figure 2.3 Uniaxial compression. Determination of Young's modulus and Poisson ratio
~--------- '---""------"-----" '" ---
..---.--- .. -------

24 Linear elasticity
Material properties 25
effect" is given by Table 2.2 Interrelations of the elastic constants of an isotropic
a= material
eyy = -V-, (2.9)
E Ev 2Gv G(E - 2G)
A --
where the Poisson ratio, V, is always positive and less than 0.5. (l + v)(l - Zv) 1 - 2v 3G-E

In general, a static deformation test consists of (1) the preparation of a specimen of E


G
prescribed form, (2) the application of stress (or displacement) at some of the bound- 2(1 + v)
aries, (3) the measurement of the resulting displacement of the boundary surface (or E 2G(1 + v)
the resulting stress on the boundary surface). The uniaxial compression test illustrated E 2G GE
K
on Figure 2.3 is suitable to determine Young's modulus and the Poisson ratio in one 3(1 - 2v) 3(1 + v)(l - 2v) 3(3G -E)
experiment. The compressive stress and the strains are readily derived according E-2G
to the expressions shown on Figure 2.3. The two material properties, E and v, are v ---2G
obtained from their definitions. E 2G 4G2
Other simple tests give rise to other material properties. During the torsion of a E' --
1- v2
--
I-I! 4G-E
circular bar around its axis, the shear stress and shear strain are related by

(2.10) where A- and G are often referred as the Lame constants. As seen, only two indepen-
dent material constants are necessary compared to the original 21.
where G is the shear modulus. Under hydrostatic compression the relative volume Thus, for an isotropic material the elastic constants E, G, v, A, K are related by
change is related to the hydrostatic pressure through the bulk compressibility, K. simple algebraic relations and any two of them determine the other ones. In the
At this point an important question arises. Is there any relation between the fracturing literature mostly E, G and v are used. Table 2.2 shows how the other
observable material properties? In other words, how many independent properties constants can be related. Some authors prefer to introduce additional combinations.
are necessary to characterize the material already known to behave linearly? Starting One of the combinations, the plane strain modulus, E', is particularly useful in
with the generalized Hooke's law: fracture mechanics, hence it is included in Table 2.2.
Static tests do not provide the only possibility to measure material properties.
a;u Cll Cl2 C13 C14 C15 C16 e;u Dynamic tests consist of periodically changing the load on the surface of the material
ayy C2l C22 C23 C24 Czs C26 Eyy and observing various characteristics of the forced elastic waves. The propagation
aZZ C3l C32 C33 C34 C35 C36 ezz
x (2.11) velocity of a longitudinal wave in the interior is, e.g., related to the density and the
axy C4l C42 C43 C44 C45 C46 Exy elastic constants according to Billington and Tate [1]:
axz C5l CS2 C53 C54 C55 CS6 exz
ayZ C61 C62 C63 C64 C65 C66 8;;z
(2.13)
where C, the stiffness matrix, consists of 36 material constants, the situation appears
cumbersome. Fortunately, the symmetry requirement, Cij = Cji, decreases this
number to 21, "which is now generally accepted to be the number of independent Since longitudinal expansion and contraction involve volume change and shear
elastic constants (Billington and Tate [1])". strain at the same time, it is not surprising that both the compressibility and the shear
Determining 21 material properties is still very difficult. Assuming some addi- modulus playa role in the final expression.
tional invariance properties, however, may further reduce the number of indepen- The great advantage of dynamic tests is that in situ measurements (without cutting
dent material constants. By far the most effective assumption is isotropy. For an out a specimen and, hence, destroying the material) are available. To characterize
isotropic material the properties are independent of direction. The stiffness matrix is the material, the velocity of two different types of waves has to be determined.
of the form
A+2G A A- D 0 D
A A+2G A-
Example 2.1 Determining Elastic Properties from a Uniaxial Test
0 0 0
). A ). +2G 0 0 0
0 0 0
(2.12) A cylindrical sandstone specimen (density, p =
2700 kg/m-) is loaded by a compressive
G 0 0 force, F = 0.8 X 106 N. The height (I = 20 em) decreases by 4 mm and the diam-
0 0 D D G 0 eter (D = 5 em) increases by 0.2 mm. Determine the elastic constants and predict the
0 0 0 0 0 G longitudinal wave propagation velocity.
---_ .. ..
' _---- ...._---_ _-----
...

Plane elasticity 27
26 Unear elasticity

Solution
The test gives Young's modulus and the Poisson ratio directly:

F 0.8 X 106
_ D2;r/4 _ 0.052 x rr/4 10
E - ---;rr- - 0.004 = 2 x 10 Pa = 20 GPa
T aT
(a) (b) (c)
so 0.2 x 10-3

v=- fl = --;2'g'ii'g04;;<5,- = 0.2.


(J~
I 0.2
From Table 2.2 we obtain

E ,
,
0
G = 2(1 + v) = 8.5 GPa, - ~
E
K = 3(1 _ 2v) = 11 GPa, (e) S

and hence, Eq. 2.13 predicts Figure 2.4 Stress-strain relations of several types of materials

VL = (3K 3p+ 4G) 1/2 _


-
(3 x 11 X 109+ 4 x 8.5 X 109) 1/2
3 x 2700
_
- 2900 m/s. 0
the envelope of stability or failure (yield) surface, A failure (or yield) criterion is the
equation of this envelope.
A detailed description of the material thus consists of an elasticity constitutive
equation, a yield criterion and another constitutive equation valid in the post-yield
2.2.2 Material Behavior Beyond Perfect Elasticity region. Since yielding is inherently irreversible, a simple curve is not enough to
represent the behavior in this region. In fact the actual behavior depends not only
A hypothetical linear elastic body "answers" with a continuously increasing strain to
on the strain but on the history of the total loading process.
a linearly growing stress. No material can be loaded infinitely, because, eventually,
it will fail. At this critical value of the stress any further "strain" can be achieved
easily because the material loses its ability to resist deformation. This is seen from
the stress-strain curve (a) of Figure 2.4. The fact that the curve (up to the failure) is a
2.3 Plane Elasticity
straight line indicates that the material is linear elastic. The dashed line represents the The description of stress and strain in three dimensions is complicated. Fortunately,
brittle rupture. The stress-strain curve (b) corresponds to a plastic material in which in many cases we can make simplifying assumptions to reduce the problem into a
strain occurs without any change in the stress. The work done in the plastic region two-dimensional one.
is dissipated and the material flows. Curve (c) shows a material with an elastic and
a plastic region. Curve (d) illustrates a material without a distinct elastic region. The
stress necessary to start any deformation is called yield stress. Curve (e) corresponds 2.3.1 Plane Stress
to a material exhibiting typical nonlinear behavior, most likely due to continuous
damage evolution. When one of the principal stresses is zero, the condition of plane stress is satisfied.
The limit of the elastic behavior depends on the type of loading. Many solids, The plane stress condition is a good approximation, for instance, for a thin plate
failing at moderate tensile stresses, may carry much higher loading in the form of (Figure 2.5)_ If the sheet is in the x, y plane, load is allowed only in the same
compressive stresses. In general, the failure will occur (or yielding will start) at plane but the deformation is not restrained in the z direction. The following stresses
specific combinations of the three principal stresses. If we represent the state of the are zero:
stress as a point in the three-dimensional stress space (taking the principal stresses as (2.14)
coordinates) then the stable and unstable states will be separated by a surface called
28 Unear elasticity Plane elasticity 29

ZtL x
au =0

r,tz'= tz:w

'YZ= 'zy =0
=0
Uyy

t tty
...

Figure 2.5 Material in the state of plane stress

It is useful to remember, however, that plane stress does not mean automatically the x
absence of strain in the z direction. (The plate may, for example "swell".)
Figure 2.6 Stresses acting within a plane

2.3.2 Stresses Relative to an Oblique Line (Force Balance I)


Now we return to the problem of determining the normal and shear stress vectors y
relative to an oblique plane. Since we deal with two dimensions (i.e, the stresses
in the x, y plane) the oblique plane is represented by a straight line. Assume that
we know the angle e and that we have a description of the stress state, i.e, the two
normal stresses (7=, Uyy and the shear stress <"xy. The definition of the stresses is Sy
shown on Figure 2.6 with the help of a square placed parallel to the axis.
In Figure 2.7 the straight line cuts off a comer of the same square. The direction
vector s and the normal vector n of the straight line are determined uniquely by
the angle e. We are interested in determining the stress vector S, i.e. calculating its
two components: Sx and Sy. Once we know the stress vector S, we also want to
decompose it into a normal component (7 and a shear component r with respect to A c
the straight line characterized by the angle e. Applying the Pythagorean theorem we
x
can write
Figure 2.7 Stresses acting on an oblique line
(2.15)

(2.16) Equation 2.17 can be rewritten as

posing restriction on the lengths of the components. Sx = Uxx case + <"xy sin e, (2.19)
Consider the balance of forces in the x direction:
and similarly, we obtain
(2.17)
Sy = <"yx cos e+ Uyy sin fJ. (2.20)
where LAB,LAC and LBC are the lengths of the corresponding line segments. Since
Since (7 is the projection of S, we have
LAB . LAC
case =- and smB =-, (2.18) (7 = Sx cos e + Sy sin e. (2.21)
LBe LBe
Linear elasticity Plane elasticity 31
30

Substituting Eqs. 2.19 and 2.20 into Eqs. 2.21, the final form of the normal stress is

(2.22)

Similarly, the shear stress is given by

r = (O"xx - axx) sin a cos a - rXy(cos2a - sin2 B)


(2.23)
= ~(O"xx - O"y)") sin(2B) - rxy cos(2a).

Equations 2.22 and 2.23 allow us to compute the normal and shear stresses relative
to a line characterized by the angle B. In Chapter 3 they will be used extensively.

2.3.3 Equilibrium Relations (Force Balance /I) Figure 2.8 Material in the state of plane strain

Up to now we have studied relations valid at a given point. Another force balance 1
puts restrictions on the variation of the stress components with the location. These eyy = E[ayy - v(O"zz + axx)], (2.28)
are the well-known equilibrium relations. For the case of plane stress (without body
forces) we have 2(1 + v) (2.29)
exy = E 'XY'

aO"xx + a,yx = 0, (2.24) with all the other stress and strain components being zero.
ax ay Remembering that the strains have been derived from displacements and in plane
arxy + Myy = o. strain only the displacements Ux and uy are not zero, it is not surprising that the three
(2.25) strain components are not independent. Indeed, their variation is constrained by the
ax ay
compatibility relation:
As we see, the equilibrium relations give two equations for the three independent
(2.30)
stress components.

These relations are automatically satisfied if we consider the displacements rather


2.3.4 Plane Strain than the strains as the unknown variables to be determined.
Plane strain is a reasonable approximation in simplified description of hydraulic
Another special case reducing the three-dimensional problem into a two-dimensional fracturing. The main question is how to select the plane. Two possibilities arise and
one is the case of plane strain. Consider a (practically) infinite body of uniform cross this has given rise to two different approaches of fracture modeling. The state of
section lying parallel to the z axis (Figure 2.8). External forces applied are parallel plane strain was assumed in horizontal plane by Khristianovitch and Zheltov [3,4)
to the x, y plane, i.e. they are "infinitely repeated" in every cross section. Intuitively and by Geertsma and de Klerk [5], while plane strain in vertical plane (normal to
it is obvious that the state of strain is independent of the coordinate z, i.e. it also the direction of fracture propagation) was assumed by Perkins and Kern [6] and
"repeats itself'. In addition, the following strains are zero: Nordgren [7). Often in the hydraulic fracturing literature the term "KGD geometry"
is used interchangeably with horizontal plane strain assumption and "PKN geometry"
(2.26) is used as a substitute for postulating plane strain in the vertical plane.
The selected plane (in which the plane strain condition is applied) should be
The strain is, thus, two-dimensional, with no displacement in the z direction. (The orthogonal to the largest extent. For a long fracture (hundreds o~ me~e~sof length)
normal stress in the z direction is, however, not necessarily zero.) The relation with limited height (tens of meters) and small width (measurable In millimeters) one
between strain and stress for the plane strain system (without body forces and temper- can assume the state of plane strain in every vertical plane orthogonal to the length
ature variation) is axes. For a short fracture (a few meters of length) with considerable height (tens of

(2.27)
meters) and small width (millimeters) one can assu~e the =
of ~lane strain ~n
every horizontal plane. In Chapter 4 the two geometnes are discussed III more detail.
.~~---.--------.---.--- ..--- ...- _.
.... __ .

32 Unear elasticity Pressurized crack 33

2.3.5 Boundary Conditions

The governing equations of any problem in linear elasticity are the same and have
to be sati~fied at all points of the body. The specific problem is distinguished from
others by its boundary conditions. In fracture mechanics often mixed boundary value c x
problems are considered where at some part of the boundary the stress and at some
other part the displacement are defined.
I ..

..'
2.4 PressurizedCrack
Linear elasticity deals with static equilibrium of solids. Fracture propagation is a
dynamic phenomenon. Nevertheless, we can still use linear elasticity to describe a
stable fracture and its neighbours. Moreover, even a propagating fracture can be
considered as transforming through equilibrium states. A "snapshot" of such an
equilibrium state will be considered in the remaining part of this chapter.

2.4.1 Solution of the Line Crack Problem Figure 2.9 Pressurizedline crack

An infinite plane with a hollow two-dimensional "crack" in it is an idealization


needed for an analytically tractable problem. In addition, the crack is assumed to be at every location. For practical purposes, however, we are only interested in the
without any appreciable opening. Pressure acts inside, trying to open the two sides displacement of the crack surface and the stress state at the tip and further along the
of the line. This is the famous Griffith [8} crack problem. We are interested in the crack axes.
(small but important) displacement of the points, i.e. in the shape of the fracture. Mathematical solutions, based on the pioneering work of Muskhelishvili [9], have
Moreover we wish to describe the state of stress around the fracture. been accomplished by solving integral equations (England and Green [10], Green and
For the sake of simplicity we assume that the infinite plane (in which the displace- Zema [11]) or applying integral transformation (Sneddon [12]).
ment occurs) is the x, y plane, and the line crack of length 2c is located along the The solution procedure starts with the construction of a function g(~) according
x axis with its center in the origin (see Figure 2.9). The half-length of the line crack, to
c, is the characteristic size associated with the problem. The boundary conditions are: t' p(x)dx
o < ~< c. (2.32)
g(~) = Jo W - x2)1/2

l7yy(X, 0) = - p(x), O::s x ::5 c,


Note that ~ is a dummy variable having the same dimension as x. The function
Uy(x, 0) = 0, X> C. (2.31) g(~) has the same dimension as the pressure and it can be considered as a modified
'1:"xy(x, 0) = 0, x::: o. pressure summing up the effect of the pressure acting not only at the given location
but at every other location. Once that function is known, the normal displacement
The first condition states that the pressure acting on the line is compensated by of any point on the upper side of the crack line is given by
the normal stress of the same magnitude. (The second argument of the unknown
functions is y. Clearly, y = 0 at the crack line.) The pressure is a function of the
uy(x. 0)
4
= - rrE'
jC
x
~g(~)d~
(x2 _ ~2)1/2' O::5x ~c. (2.33)
location x. It is supposed that the problem is symmetric with symmetry axes x = 0
(the pressure acts on both faces) and y = 0 (the function p(x) is an even function.)
Since the above boundary conditions are written for the upper right quadrant only, where E' is the plane strain modulus, given in terms of the other properties in the
the requirement for p(x) to be an even function is not restrictive in this half-plane, last row of Table 2.2.
but it is understood that p( -x) is defined to be equal to p(x). By virtue of the second boundary condition, the displacement is zero outside
It is assumed that all stresses disappear in infinity, i.e. the far-field stress state is the crack. Clearly, a similar displacement of the lower line occurs in the negative
zero. Non-zero far-field stresses will be treated in Section 2.6. The complete solution y direction, and hence the width of the crack is simply twice the value given by
of the problem includes the construction of the stress state and the displacements Eq. 2.33. The normal stress in the y direction is known inside the crack (it is the
34 Linear elasticity Pressurized crack 35

opposite of the pressure). Along the crack axes (but outside the open interval) the
normal stress is given by

2 [ xg(c) r
g'(~) dS-- ]
O"yy(x, 0) = -; (X2 _ C2)1/2 - g(O) - x Jo (X2 _ ~2)1/2 ' x c. (2.34)

The above equation is the one given by Sneddon (12, p. 318] with a modification. It
is derived for the case of differentiable g(~). If the pressure is a continuous function c c
of the location, then g(~) is differentiable. Fortunately, Eq. 2.34 gives correct results x
even for those cases where there are some distinct jumps in the pressure. Along
the line y = 0 the normal stress in the x direction equals that in the y direction:
O"=(x, 0) :;; O"yy(x, 0), X > c. Along the same line the shear stress, 't'xy(x, 0), disappears
(see e.g. Green and Zema [11, p. 276]).
For several specific p(x) functions of practical importance, the above integrals
can be solved in closed form.
b

2.4.2 Constant Pressure


c c 2c 3c 4c
If the pressure opening the crack is constant, Po, the g(~) function is given by x x
($ podx porr Flgure 2.10 Constant pressure solution to the line crack problem
g(~) = l (~2 _x2)1/2 = T' o < ~< c. (2.35)
While the shape of a pressurized crack is easily treatable, the stress distribution
Therefore, the corresponding g(~) function is also constant. The resulting displace-
is more difficult to handle. The last term in Eq. 2.34 is zero since the function g(~)
ment of the upper line of the crack is is constant. The remaining terms yield

Uy(X, 0)
4
= - E'rr
Ie
x
~d~
(x2 _ ~2)1/2 =
2po f"22.
Ii! V c: - .r-, o:::::x::::: c, (2.36)
O"yy(x,O) = Po [(X2 :c2)1/2 -lJ, x> c (2.39)

This is the well-known result stating that the line crack is of elliptical shape with and hence the stress is infinite at x = c, i.e. at the crack tip. The phenomenon is
the width given by often referred to as stress singularity.
w(x) = 4po
-y
E'
~
c2 -x2 (2.37) The solution of the constant-pressure line crack problem is illustrated by
Figure 2.10 consisting of four plots. The first plot gives the pressure acting along
The width is zero at the end of the crack (at x = c) and the maximum width the crack line. The second is the function g(~). It represents the integrated influence
occurs at the middle of the crack (at x == 0): of the pressure at a given location ~. The third plot shows the displacement (in other
words the half-width,) u(x). Note the characteristic ellipsoid shape. Finally, from the
4cpo plot of the tensile stress ahead of the crack tip, the dramatic effect of the crack can
WO=P' (2.38)
be well seen. The stress induced by the inner pressure and amplified by the crack
is infinite at the tip but decreases rapidly with the distance from the tip. The effect
The reader may feel uncomfortable that the crack is treated as a line on the one
of the crack is hardly observable at distances larger than a few multiples of the
hand and has an elliptical shape on the other. The given approach can be justified
crack length.
only if the width is orders of magnitude smaller than the length. Fortunately this is
the case for hydraulically induced fractures.
Perhaps the most interesting feature of the solution is that the linear behavior 2.4.3 Polynomial Pressure Distribution
valid locally is preserved for the whole crack as an entity. Indeed, the width depends
linearly on the opening pressure. The proportionality factor between pressure and Interestingly, very few solutions have been published for any pressure distribu-
width contains the elastic properties and a characteristic length. tion different from constant. Here we give (some new) results for a polynomial
36 Linear elasticity Pressurized crack 37

distribution 0.5 0.8


(2.40) 0.4 ~0.6
t~ 0.3 ro
a,
e. 0.4
Since any pressure distribution can be approximated by its Taylor series, the knowl- Q. 0.2 c:"
edge of the solution is of great importance. 0.2
0.1
Interestingly, the g(~) function is a polynomial in ;:
0 0
0 2 4 6 8 10 0 2 4 6 8 10
x(m) ~(m)
(2.41)
8 1.2
The normal displacement of the upper crack line is given by
6
<? 0.8
E c..
uy(x, 0) = ~,{ 2po(c2 _ x2)1/2
..4
~
e..
'" 0.4

+ 2:1 [CCc2 _ x2)1/2 + x2 In C + (c2x- X2)1/2]


2

0 0
-,
0 2 4 6 8 10 10 20 30 40
x(m) x(m)
2
+ f_(cZ + 2x2)(? _x2)1/2 Figure 2.11 Linear pressure drop solution (Example 2.2)
3

+ .~ [(~c3 + cx2) Ccl _ xZ)lf2 + x4ln c + CCZ-: X2)1/Z] + ... } (2.42) Solution
Because of the horizontal plane strain we can use c XI = =
10 m. The pressure gradient
and the normal stress distribution is
is PI = -PO/xI =
-50 kPaim while P2 = =
P3 o. Figure 2.11 shows the important
characteristics of the fracture. The width is twice the normal displacement:

CTyy(X, 0) = po[x(x2 - cZ)-1/2 - 1]

+ 2Pl [cx(x2 _ C2)-1/2 _ x arctan(c(x2 _ c2)-1/2)]


l( The maximum width is located at the center of the fracture, i.e. at x = O. If x tends to
zero, the second term diminishes and the width is equal to
+ P2 [ ( c~x _ ~ ) (x2 _ c2)-1/2 _ x2] 4
Wo = -(poJr+ PIC)C.
E'Jr
Using the specified data
+ 2:3 [( ~x -cx3) (x2_c2rl/2_~ arctan(c(x2-C2)-1/Z)] + .... (2.43)

Wo = 4{1-109]!"
0.252)
(5 X lOSJr - 5 X 104 X to) x 10 = 12.8 X 10-3 m = 12.8 mm.
The reader may easily verify that the above expressions reduce to the constant-
I! is interesting to compare the characteristics of the solution (Figure 2.11) with
pressure case, discussed previously, if the higher-order coefficients are zero.
those shown in Figure 2.10. The displacement, u(.x) has a very similar profile in both
cases. Also the shape of the stress distribution, u(x), is very similar near the tip in the
two cases. 0
Example 2.2 Line Crack with Linear Pressure Drop
In a fracture of half-length XI =
10 m the pressure is Po = 0.5 MPa in the center and 2.4.4 "Zipper" Cracks
linearly drops toward the tips. At the tip the pressure is zero. Assume the state of
plane strain in (any) horizontal plane. What is the maximum width if E = 1 GPa and The distributed forces, denoted by p(x), might be of cohesive nature as well. There-
v = O.25? fore, we can allow negative p(x), at least in some part of the fracture. This "negative
--- ~--. "-_ ... _-_ ....~-.--. --_ ... -_. __ ...

38 Unear elasticity Pressurized crack 39

pressure" tries to close the fracture and may result in exotic fracture shapes. Of 1.5 1.8
particular interest is a shape with smooth closing at the tip because then the stress 1 1.5
singularity disappears. Barenblatt [13J, who has built a whole theory on the condi- 0.5
0 -------------------- <0 1.2
tion of smooth closing, notes that one of his colleagues compared such fractures '"
o,
-0.5 D.
0.9
to zippers. ~ -1 ~
o_ C 0.6
A glance at Eq. 2.34 may convince us that for avoiding stress singularity at the -1.5
-2 - 0.3
tip, the function g(~) should be zero at ~ =: c. This is the zipper crack equation. If -2.5 0
the function g(~) is zero at the tip, the displacement computed from Eq. 2.33 has a 0 2 4 6 8 10 0 2 4 6 8 10
zero derivative at the tip with respect to x and hence the crack is closed smoothly. x(m) ~(m)
A crack opened by constant positive pressure cannot have smooth closing. If, 2
however, we allow for two levels of the pressure, one positive near the center of the 0.6
crack and one negative near the tip, the condition of smooth closing can be satisfied. 1.5
This special pressure distribution plays a key role in several studies including those of E <0
c, 0.4
..
Khristianovitch and Zheltov [3,4], Barenblatt [13] and Geertsma and de Klerk [5]. ::> ~
For brevity we refer to that pressure distribution as the jump function. The jump 0.5 b 0.2
function will be characterized by three constants: Sl is the repulsive distributed force
(pressure) in the center part of the crack, 4 8 10 20 ~o 40
x (m) x (m)
p(x) =: S1, X ~ Xo, (2.44)
Figure 2.12 Zipper crack solution (Example 2.3)
and S2 is the cohesive distributed force (negative pressure) near the crack tip:
if x ~ xo,
p(x) = -S2, xo < x ~ c, (2.45)
if x> Xo.
The third parameter is the location of the jump, Xo (see Figure 2.12). Note that all
the parameters are positive. The corresponding g(~) function is given by The maximum width is twice the normal displacement at x = 0, i.e.
8
Wo = Ire (51 + sz)xo
(C
In ~ + q1)
x Xo (2.46)
1+ cos ( res2 )
=:
8(SI + s2)e sm
. [ 1(S2 ] x In 2(51 + S2) (2.49)
and the zipper crack equation, gee) = 0, is satisfied if E'tt 2(SI + 52) . (
sm resz )
1(S2 2(SI + S2)
Xo = esm . (2.47)
2(51 + S2) Final!y, the stress at the tip is
This condition was first derived by Khristianovich and Zheltov [3]. The corre-
sponding displacement is obtained from Eq. 2.33. Again, two different expressions 2 [ x(e2 - xt3)1/2]
should be used, depending on the location x. O'yy(x,O) =: -SI + - (Sl + S2) arctan 2 2 1/2 . (2.50)
1( xO(X - xo)

. 1
uy(x,0)=-,-(Sl+S2)
[ql
4xoln---+xln
+ q2 e2x5 - 2qlq2XQX
22
+ ?x2 - 2x5X2]
E 1( q3 C Xo + 2qlq2xoX + C22X - 2xOX
22'
Example 2.3 Zipper Crack with Piecewise Constant Pressure
where(2.48)
=
Assume that 51 1 MPa pressure opens a line crack of half-length c = 10 m and the
ql = (e2 - X5)1/2, distributed force (negative pressure), 52 = 2 MPa tries to close it near the tip. Where
is the jump of the pressure located if the ending is smooth? Calculate the maximum
q2 = (c2 - x2)1/2, =
width if Young's modulus is E 9.1 GPa and the Poisson ratio is v 0.3. =
40 Linear elasticity Stress concentration and stress intensity factor 41

Solution 0.5 ~--------.


The location of the jump is calculated from Eq. 2.47. 0.3 0.6
r0-
~
.:to = 10 x
. 1l' X 2
SID ---
r:
= 10",,3/2 = 8.66 m. e. 0.1 a,
e 0.4
2(1 + 2) Q.. 0>
-0.1 0.2
The plane strain modulus is
-0.3 o ~~~~--~~~~
0 2 4 6 8 10 o 2 4 6 8 10
E' = 9.1 X 109 = 1010 Pa x(m) x(m)
1 - 0.32 '

and the maximum width according to Eq. 2.49 is given by 6 0.25 r---------,
0.2
8 x (3 106) x 8.66 ( 1 + 0.5) E'4
woo = X
10
10 it
In ---
0.866
= 3.63 X 10-3 m
.
= 3.63 mm. lO.15
. :i
::. 2 -; 0.1
Figure 2.12 illustrates the solution. Note the smooth ending of the fracture. The normal
stress is finite at the tip and equals to the normal stress acting on the fracture face inside 0.05
the tip (2 MPa) but it still has negative infinite derivative. 0 0 o L-~""""':::::::==d
0 2 4 6 8 10 10 20 30 40
x (m) x(m)

2.4.5 "Zipper" Crack with Polynomial Pressure Distribution Figure 2.13 Zipper crack solution with linear pressure drop (Example 2.4)

Applying the zipper crack equation to the g(~) function given by Eq. 2.41, the
The "pressure" at the tip is
following simple condition is obtained:

3npo + 6CPl + 3nc2 p~ + 4c3 p~ = O. (2.51)


Po - PIC =- PoJr
2c
= 0.5 X 106 Pa - (78.5 x 103 Pa/m)(10 m)

Any pressure distribution in the form of a cubic polynomial and satisfying Eq. 2.51 = -0.285 X 106 Pa = -0.285 MPa.
results in smooth closing. Some care should be taken to ensure, however, that the Similarly to Example 2.2 the maximum width is given by
displacement should be nonnegative. (Otherwise the crack is "overdosed" and the
boundary conditions are no longer valid.) 4
The corresponding displacement and stress distribution can be read directly from
Wo = -(PoJr
E'rr
+ PIC)C.
the equations given in Section 2.4.3. Substituting the given parameters, we obtain

4 x (1 - 0.252)
Example 2.4 Linear Pressure Drop Causing Smooth Closing Wo = 09
1 x rr
(5 X 1<f1l'- 7.85 X 104) x 10 == 9.37 X 10-3 m = 9.37 mm,

In a fracture of half-length XI = 10 m the pressure is po = 0.5 MPa in the center and Figure 2.13 illustrates the solution. Note the smooth closing at the tip. The normal
drops linearly toward the tips. Assume the state of plane strain in (any) horizontal plane stress at the tip is Po(-l +1l'/2) =
0.285 MPa. This is the same value as inside the
considering a line crack with C = xI. What is the "pressure" at the tip if the fracture crack. The normal stress declines with the distance from the tip rather similarly to
closes smoothly? What is the maximum width if E =1 GPa and v = 0.25? (Note the Example 2.3. 0
similarity and difference with respect to Example 2.2.)

2.5 Stress Concentration and Stress Intensity Factor


Solution
Fracture mechanics has emerged from the observation that any existing disconti-
After substitution of P2 = P3 =0 into Eq. 2.51 the pressure gradient satisfying the nuity in a solid deteriorates its ability to carry loads (see Griffith [8J; Inglis [14]).
zipper equation is A (possible small) hole may give rise to high local stresses compared to the ones
PI = --Po1l'
2c
= -78.54 kPa/m. being present without the hole. The high stresses, even if they are limited to a small
----------------------

42 Linear elasticity Fracture shape in the presence of far-field stress 43

area, may lead to the rupture of the material. It is often convenient to look at mate- PoT( .
rial discontinuities as stress concentrators, especially if the nominal stress state (the
gee) = '2 + PIC = 2.85 X 10' Pa. (2.54)

"would be" stress state without a hole) can be scribed by a single scalar. Thus from Eq. 2.53 the stress intensity factor is
A circular hole in a plane induces stresses, the maximum magnitude of which is
UP x 101/2
three times the far-field stress. In other words a circular crack concentrates the stress
and the stress concentration factor in this case equals 3.
K/ = 2g(e)cI/2
7r
= 2 x 2.85 x l'(
= 5.7':J_ X 105 Pa m1/2. 0

The stress concentration factor is insensitive to the actual size of the crack. The
nature of the far-field stress is also of limited importance. What really matters is
the shape of the crack. If the radius of the crack in the x direction is considerably 2.5.2 Stress Intensity Factor, Non-symmetrtc Loading
smaller than the radius in the y direction, i.e. the crack has an elliptical shape, the
Equation 2.53 was derived for a crack with symmetric loading, i.e. for p(x) = p( -x).
stress concentration factor may be very large. Inglis [14] showed that in the limiting
If the loading is asymmetric, there are two different stress intensity factors at the
case of a sharp crack the stress concentration factor tends to infinity since the stress
two tips (see Rice [15]):
at the tip becomes singular.
In fracture mechanics we have to deal with singularities. Two different loadings
(pressure distributions) of a line crack result in two different stress distributions. Kl+ = 1
r.;:;:
yl'CC
jC -c
p(x) mx+x dx,
---
c-x
Both cases may yield infinite stresses at the tip, but the "level of infinity", or simply
the "amplitude" (see Sneddon [12]) is different. We need a quantity to characterize
this difference.
K/_ = 1
r.;:;:
yl'CC
t -c
p(x) ~-x
--
c +x dx.
(2.55)

Equation 2.34 shows that the singularity is always of the same type. In fact the
singular stress multiplied by rl/2, where r is the distance from the tip, will be always The above relations are of interest for determining the height of a fracture as will be
finite. Therefore the stress intensity factor K/ is defined as discussed in Chapter 11.

(2.52)
2.6 Fracture Shape in the Presence of Far-field Stress.
Not surprisingly, the stress intensity factor is strongly related to the function g(;), The Concept of Net Pressure
introduced in Eq. 2.32.
Up to now we have considered a pressurized line crack in the absence of far-field
stresses. How can we apply the results to real fractures where the far-field stresses
2.5.1 Stress Intensity Factor, Symmetric Loading are not zero? The general answer to this question is fairly complicated. Fortunately a
larger fracture will align itself to the far-field stress field as discussed in Chapters 3
The stress intensity factor of a pressurized line crack is readily derived from Eq. 2.34: and 4. Therefore, as a first approximation, we can represent the fracture by a line
2g(c)c1/2
crack located parallel to the-maximum principal stress of a given plane and opening
KJ=--- (2.53) up in the direction of the minimum principal stress of the same plane. The stresses
T(
are undisturbed only (relatively) far from the existing fracture and hence we call
For a constant pressure line crack KJ = pocI/2. If the pressure distribution results in them far-field stresses.
a zipper crack then the stress intensity factor is zero. For consistency with the previous sections it is convenient to direct the x axis in
the direction of the maximum principal stress. Figure 2.14 shows the configuration
of the line crack and the definition. We assume that at infinity the normal stresses
Example 2.5 Stress Intensity Factor are constant, satisfying ay < ax, and the shear stress is zero, Txy O. =
For this configuration the principle of superposition applies [2] as shown on
Computethe stress intensity factor for Example 2.2.
Figure 2.14. In practice this means that an associated problem can be created. The
associated problem has the same structure as the problem defined in Section 2.4.1.
Solution The only difference is that at the crack boundary the algebraic sum of two forces
has to be considered. The algebraic sum is the difference of the inner pressure and
Substituting c = 10 m, Po = 500 kPa, PI = -Po/X! = -50 kPa/m into Eq, 2.41, the the far-field minimum principal stress. It plays a central role in hydraulic fracturing.
end value of the g(~) function is obtained as It is called net pressure.
Circular crack 45
44 Unear elasticity

and
y y y if x ~xo
if x> Xo
X x
---l--_'
a"y Substituting E' == 10lD Pa, 51 == 106 Pa, S2= 2 X 106 Pa and xf == 10 m, the width is
obtained as a function of the location x:

p 6 10-4 [ 5 + ,JI00 - x2 150- x2 - xj3(100 - X2)]


p-ay x 20.J31 +xln '
w(x) = n n .J75 _ x2 150 _ x2 + x.j3(100 - x2)

= + ------- ------_. o <x < 8.66 m

Original problem Associated problem Trivial problem 6- x 1~-4 [20.J3 I 5 + ,J100 - xi 150 - xl - xJ3(100
+ x In 150-x2+xJ3(100-x2)
- X2)]
'
w(x)== n n .Jx2-75
Figure 2.14 Pressurizedcrackin the presenceof far-fieldstress.The principleof superposition
8.66 < x ~ 10 m.
The solution of the associated problem is then readily transformed to yield the
solution of the original problem. At x _ 0 the width is Wo == 3.63 mm as we know alrea~y from Example 2.3.. The
The principle of superposition will be illustrated in conjunction with Example 2.3. normal stress distributionO"yy(x, 0) is essentiallythe same as m Example 2.3, but shifted
The notations of the classical paper of Geertsma and de Klerk [5] will be used. with the constant far-fieldvalue, (J" yo _. .
Equation A.2 of the paper by Geertsma and de Klerk [5] gIves the WIdthm the
Example 2.6 Zipper Crack in the Presence of Far-Field Stress following form:

Assume that the inner pressure, P = 3 MPa opens a line fracture of half length x f =
1 _2 2) 1/2
10 m against a far-field compressivestress Smi. = cry = 2 MPa. The exact magnitude
of the other principal stress is not known, but cry < a, and 'xy == O. Near the tips,
4 _ 1 - "7
XOq4!
w(x) == -E'Ir p x In 1+ XOQ4
[ __
! -
11 - q41
Xo In 1 + Q 4 where Q4 = (-2--2
Xf -x
Xf - Xo

Xo < Ixl ~ XL, the pressure is virtually zero (i.e. negligible). Assumingplane strain in X
the X, y plane find the location Xo if the fracture ends smoothly. Give the width as
a function of the location. Young's modulus is E = 9.1 GPa and the Poisson ratio It is easily found numericallyor by doing some algebraic manipulationsthat the two
is v = 0.3. expressions give identical result. ." . .
Geertsma and de Klerk [5) also give the following approxImationto wo, which IS
Solution valid if Xo tends to xf:
8 _XO~2 Xf -xo
Wo = -p-
By virtue of the superpositionprinciple we can use the solution of Example 2.2 with
IrE' xf
SI = P - Oy = 1 MPa and Sz = (J"y == 2 MPa and half length c == xf == 10 m. The loca-
tion Xo is exactly the same as in Example 2.3:
For this example the above expression gives Wo == 3.3~ ~ (10~ e;,ror).The approx-
Xo == 10 x sin [ 1T x 52 ] = 10 x sin [ Ir x2 ] = 8.66 m. imation improves rapidly if the ratio plO"y tends to unity, draggmg Xo toward xf' 0
2(sl + S2) 2x (1 + 2)

The fracture width is obtained from Eq, 2.48, multiplying the normal displacements
by 2. 2.7 Circular Crack
The mathematical treatment of the pressurized circular crack problem is si~nj~arto
the line crack problem. For the sake of simplicity we assume that the crac~ IS l~ the
where x y plane, opening up in the z direction (Figure 2.15). The crack surface IS a ~I.rcle
ql = (x} - X~)l/2,
located around the origin. The far-field stresses are zero. The boundary conditions
at z = 0 are
q2 = (x} _~)1/2,

-~----------- ..-
----_. __ ._-.
_. - - """'--- _._---------_ ..,----------:-- ------- .. _----"" - .._-,,_ ..... _-

46 Linear elasticity Volume and strain energy 47


y
resulting in the maximum width

8Rpo
wO=-- (2.61)
rrE'
Other cases may be treated similarly to the line crack problem.
x

2.8 Volume and Strain Energy


In this book every extensive quantity associated with the fracture is reported for one
wing of a two-wing fracture. This convention is set to ease the further use of the
formulas in hydraulic fracture models.
x The elastic energy or, in other words, the strain energy is the excess energy stored
in the medium due to the presence of the fracture. For a line crack in plane strain
condition a slice of the elastic medium of thickness 8 is considered. The thickness
Figure 2.15 Pressurizedcircular crack
is necessary; otherwise, the line crack is only a two-dimensional abstraction. The
strain energy is calculated from the definition of work: force multiplied by path.
azz = -p(r), 0 ~ r ~ R, The integral
Uz = 0, r > R, (2.56) Wo = !8foe Pn (.x)w(X) dx (2.62)
axz = 0, r;::: O.
is the work done while opening one wing of a two-wing "line crack" with half length
c, if the far-field stress is zero. The factor 1/2 is necessary because the force to open
The first condition states that the pressure acting on the surface is compensated by
the crack at any location x varies linearly with the displacement. The pressure Pn (x)
the normal stress of the same magnitude. The pressure is a function of the distance
is only the final value. Sneddon [12] shows that an equivalent expression in terms
from the origin, r, It is supposed that the problem is symmetric with symmetry
of g(~) is
plane z = O.
The first step of the solution is again the construction of a function, G(~): Wo= 8~
rrE'io
r ~[g(~)fd~. (2.63)

G _
. (~) - rr
2 r
io
rp(~)d,
(~2 - r2)1/2' o < ~< R. (2.57)
For constant pressure this reduces to

rrP5c2 (2.64)
Wo=8--.
Once the G(~) function is known, the normal displacement of any point on the upper 2E'
side of the crack disk is given by The corresponding expression for one wing (i.e. one half) of a circular crack is

o~ r ~ R. (2.58) Wo= ~ t[G(~)fd~, (2.65)


rrE' )0
and for constant pressure this reduces to
For a "zipper" fracture G(R) should be equal to zero.
The crack width is the normal displacement multiplied by 2. In particular, the
constant pressure case is characterized by (Green and Zerna [11]): (2.66)

(2.59) Notice that here we do not have to consider a slice of thickness 8, because a circular
G(~) = po~
crack is a three-dimensional object in contrast to the two-dimensional line crack.
w(r) = ~ POv'R2 - ,2 (2.60) If the far-field stress is not zero, an additional term, Wi, appears in the elastic
rrE' energy. It is readily derived from the definition of work as the product of the minimum
--- --- .._--- --_. ------

48 Linear elasticity Computational methods 49

principal stress and the volume of the fracture The total elastic energy is

(2.67) 2W == 2Wo + 2WI = 5.66 x laJ + 1.13 x 105 = 1.19.x 105 J.

Here we consider the volume of a fracture opened by constant net pressure. For The strain energy due to the net pressure for a circular crack is (from Eq. 2.66):
a line crack of half-length c the volume of one wing located in a slice of thickness
8p2 R3 8 x (3 X 106)2 x 13
8 is obtained from 2Wo = _n_O_ =
3' 3 x 1010
= 2400 I.
v=8 r 4Po
Jo E'
J c2 _ x2 dx = rr:Pno8C2
E'
(2.68) The volume of a circular crack can be determined as

16 XR3 PnO 16 x (1)3(3 x 106) 6 -3 3


As a by-product we obtained the average width, which is the volume divided by the 2V = 3E' = 3 x 1010 = 1. x 10 m.
area, i.e
iii = rr:Pno8c2 _ rr:P"oC The superposed strain energy is
(2.69)
E'8c - E'
2WI = 2VUmin = 1.6 X 10-3 x 3 X 107 = 4.8 X 104 J.
The volume of one wing (one half) of a circular crack is obtained from
The total elastic energy of the circular crack is
V = -21 rR (2m)
Jo
(8 P"o .; R2 _
E'
r2) dr = 8R3 PnO ,
3E'
(2.70) 2W = 2Wo + 2W I = 2.4 x laJ + 4.8 x 1(f = 5.04 X 104 J. 0

indicating an average width, During the fracturing process we have to provide the elastic energy to open the
fracture and an additional energy that is dissipated (i.e. heats the formation). As
8R3 PnO 16PnoR usual, the second. term is less known and depends not only on the final fracture
w---""':"".......,_=--- (2.71)
3rr:E' shape but on the history of creating it.
- 3E' (R~7r:)

2.9 Computational Methods


Example 2.7 Energy of Constant Pressure Fractures Finite difference, finite element and boundary element methods are available to solve
Calculate the elastic energy necessary to open two wings of a line crack of half- the linear elasticity problem with specified boundary conditions. Finite differences,
=
length 1 m contained in a 0 2 m thick slice of an infinite medium of plane strain the basic tool of solving differential equations on domains of simple geometry, have
modulus, E' =10 GPa. Compare it with the energy of a circular crack of radius 1 m. only limited application in elasticity calculations.
=
The minimum principal stress is Umin 30 MPa. The opening pressure is Po 33 MPa = In finite element methods (see Hilton and Shih [16]) the region surrounding the
(i.e. the net pressure is 3 MPa). fracture is divided into a network of elements. The "solution" consists of finding
values of displacements and stresses at the mesh points of the network. The solution
between the nodes is expressed in a simplified way in terms of the node values. A
Solution sufficient number of algebraic equations is deduced relating the approximation to the
original partial differential equation. The resulting system of linear equations is large
From Eq. 2.64 the strain energy due to the net pressure is
but sparse.
2Wo = 0__
rr:p2c2
= 2 rr: x (3 x 106 i x 12
= 5660 J.
In boundary element methods (see Crouch and Starfield [17]) only the boundary
E' 1010 is discretized. The numerical solution is sought as a linear combination of the
The volume of the two-wing fracture is known analytical solution of simple boundary value problems. The method effec-
tively reduces the dimension of the problem by one. The resulting system of linear
equations is much smaller than in the finite element method but it is dense.
Whatever method is used, the singularity of the solution requires sophisticated
The superposed elastic energy (work done against the minimum principal stress) is approaches. The intuitive answer to the challenge is to use uneven mesh, high element
concentration near the sharp edges and crack tips. Unfortunately, serious questions
2Wj = 2Vumin = 3.77 X 10-3 x 3 X 107 = 1.13 x 105 J. arise concerning the reliability, convergence and convenience of this approach.
50 Unear elasticity References 51

For a given finite element, calculation improvement on the accuracy can be 16. Hilton, P.O. and Shih, G.c.: Applicationsof the Finite Element Method to the calcu-
achieved by retaining higher-order terms in the asymptotic expansion of the lations of Stress Intensity Factors, in Mechanics of fracture I, Methods of analysis and
solutions of crack problems, Sih, G.C (ed.), Nordhoff International,Leyden, 1973.
displacement components about the crack tip [16]. In boundary value methods the
17. Crouch, S.L. and Starfield,AM.: Boundary Element Methods in Solid Mechanics,
accuracy is improved by applying special crack tip boundary elements [17]. In
Unwin Hyman, London, 1990.
both cases previous knowledge on the shape of the analytical solution to simplified 18. Clifton, R.J. and Abou-Sayed,AS.: On the Computation of the Three-Dimensional
problems is built into the solution algorithm and the "numerical solution" is in fact Geometry of HydraulicFractures, paper SPE 7943, 1979.
a combination of numerical and analytical techniques.
Solving the elasticity problem at specified loading conditions is still only a part of
the solution of the fracturing problem itself. The main difficulty is still ahead: One
has to deal with the propagation phenomenon (see Clifton and Abou-Sayed [18]).
This will be the subject of subsequent chapters.

References

1. Billington,E.W. and Tate, A: The Physics of Deformation and Flow, McGraw Hill,
New York, 1981.
2. Fenner, R.T.: Engineering Elasticity, Application of Numerical and Analytical Tech-
niques, Ellis Horwood,Chichester, 1986.
3. Khristianovitch,S.A. and Zheltov, Y.P.: Formation of Vertical Fractures by Means of
Highly Viscous Fluids, Proc. World Pet. Cong., Rome, 2, 579, 1955.
4. Zheltov, Y.P. and Khristianovitch,S.A: On the Mechanismof Hydraulic Fracture of an
Oil-BearingStratum, Izv. AN SSSR, OTN, (No.5), 3-41, 1955.
5. Geertsma,l. and de Klerk, F.: A Rapid Method of Predicting Width and Extent of
HydraulicallyInduced Fractures, JPT 1571-81, (Dec.), 1969.
6. Perkins, T.K. and Kern, L.R.: Width of HydraulicFractures,JPT 937-49, (Sept.), 1961;
Trans. AlME, 222.
7. Nordgren, R.P.: Propagation of a Vertical Hydraulic Fracture, SPEJ 306-14, (Aug.),
1972; Trans., AlME, 253.
8. Griffith,AA: The Phenomenonof Rupture and Flow in Solids, Phil. Trans. Roy. Soc.
A 221,163-198,1920.
9. Muskheiishvili,N.I.: Some Basic Problems of the Mathematical Theory of Elasticity,
Nordhoof, Holland, 1953.
10. England, AH. and Green, AE.: Some Two-dimensionalPunch and Crack Problemsion
Classical Elasticity, Proc. Cambridge Phil. Soc., London, 59, 489-500, 1963.
11. Green, AE. and Zema, W.: Theoretical Elasticity, Oxford University Press, London,
1968.
12. Sneddon, LN.:IntegralTransform Methods,in Mechanics of fracture I, Methods of anal-
ysis and solutions of crack problems, Ed. Sih, G.C. Nordhoff International, Leyden,
1973.
13. Barenblatt, G.I.: Mathematical Theory of Equilibrium Cracks, in Advances in Applied
Mechanics, 7,55-129, 1962.
14. Inglis, C.E., Stresses in a Plate due to Presence of Cracks and Sharp Corners, Trans.
Inst. Nav. Arch. 55, 219-230, 1913.
15. Rice, J.R.: Mathematical Analysis in the Mechanics of Fracture, in Fracture,
Leibovitz,H. (ed.), Vol.2, Academic Press, New York, 1968.
_. __ --,----------------

3
STRESSES IN FORMATIONS

3.1 Basic Concepts


Formations that are candidates for hydraulic fracturing are often at great depth,
overlain by earth formations of considerable thickness. This "overburden" causes
the formation of interest to be subjected to stresses. These stresses are caused
not only from geological depositional mechanisms but also from historical tectonic
phenomena.
As shown in Chapter 2, stresses are vectors. Away from a well they are often
referred to as far-field stresses, The drilling of a well creates a different near-well
stress concentration which may be much more complicated than the far-field stresses.
In considering a randomly oriented plane of area D.A at a point P within a body,
and across which a resultant force b..F acts (Figure 3.1), the stress at the point is
defined as
a = lim (b..F) ,
M ......O M
(3.1).

Since an infinite amount of planes can be drawn through a given point and although
the resultant force acting on these planes is the same, the stresses acting on them are
different because of the varying inclinations of the individual planes. For a complete
description of the stress field it is necessary to specify not only the acting force
magnitude and direction, but also the surface upon which the force acts.
Stress, applied on a surface at any angIe, can be decomposed into three vectors, a
normal stress, ax (along the x axis and normal to the x plane) and two shear stresses,
Tyx (along the y axis) and TIS (along the z axis). In a three-dimensional body the
stress tensor was given in Chapter 2 by Eq. 2.3.
Suppose that in a two-dimensional system as shown in Figure 3.2, ax and ay are
the normal stresses on the y and x planes respectively, and rxy = Tyx are the shear
stresses. Any other direction, at angle e from the y axis, would result in a normal
and a shear stress (See Roegiers (1]). These were shown in Chapter 2 to be

(3.2)

and
(3.3)
54 Stresses in formations Stresses at depth 55

3.2 Stresses at Depth


A formation at depth H can be considered as a system subjected to three principal
stresses, one vertical and two horizontal. These are also the far-field stresses. An
analysis of the effect of these stresses in hydraulic fracturing was introduced by
Hubbert and Willis [2]-
Tzx
The easiest to understand is the absolute vertical stress, O"y, which is simply the
weight of the overburden

a; = g 1" pdH, (3.6)

-ryx where p is the density of the overlaying strata. Density logs can readily provide the
density values from the surface to the formation of interest and integration of the
Figure 3.1 Force and resulting normal and shear stresses on a surface density log provides a.:
Typical values of rock density range from 2500 to 2750 kglm3 and, assuming that
all overburden consists of sandstone with p = 2650 kg/nr', a common first approxi-
y
mation of the gradient of the overburden stress is simply

dO'. 4
tyx - = (9.8)(2650) = 2.6 x 10 (palm). (3.7)

~i Y or
dH

A well-known value in petroleum field units is 2.49 x 104 Palm (1.1 psi/ft).
The calculated stress and stress gradient from Eqs. 3.6 and 3.7 are absolute, and
in a porous medium they are carried by both grains and the pore-inhabiting fluid.
0 x Biot [3] has introduced the poroelastic constant, a, such that an effective stress, a',
--<xy
on the grains is
uyt (3.8)
Figure 3.2 Normal and shear stresses acting at an angle q from a direction with known stress
values where p is the pore pressure.
It is important here to realize that the overburden absolute stress remains constant
throughout the time of interest whereas the effective stress may change profoundly
From Eqs. 3.2 and 3.3 it can be concluded that there exists an angle e where the with fluid production or injection and the associated reservoir pressure changes. Also,
shear stress vanishes and, thus, r =: O. From Eq, 3.3, setting it equal to zero, overpressured formations may have much smaller effective stresses than underpres-
sured ones.
The Poisson ratio, V, is the ratio of the lateral strain to the longitudinal strain,
o _- 21tan -1 ( ---,2-cyx )
(3.4)
O"x - O"y
(3.9)
and by substitution in Eq. 3.2
From the general elastic stress/strain relationship, using Young's modulus, E,

(3.5) 0" = Ee, (3.10)

and in a three-dimensional porous medium the following can be written


The stresses calculated from Eq. 3.5 are referred to as principal stresses and
coincide with the directions where shear stresses vanish. All other directions have
nonvanishing shear stress components. (3.11)

------_ ....--- .
56 Stresses in formations
Stresses at depth 57

If the lateral movement is constrained, i.e. Cx = 0, if a; - O"y (perfect isotropy in From Eq. 3.5
the horizontal plane) and O"r = O"y, then
O"max.min = ~(3.5 X 107 + 4.1 X 107) )(7 x 106)2 + ~(3.5 X 107 - 4.2 X 107)2
! f f V f
O"h = 0" = O"y = --O"v'
1- V
(3.12) = 4.58 10 and 3.01 x 10' Pa.
X 7 0

Substitution of the effective by the absolute stresses leads to


Example 3.2 Stresses vs. Depth
V
O"h = --(O"v
1- v
- aP) + aP. (3.13) 1. If a reservoir is 3000 m deep, the poroelastic constant 0.72, the formation a=
density Pr =
2650 kg/nr', the oil density Po = 850 kg/nr', and all saturating fluid is
This stress is the minimum horizontal stress deriving from the Poisson ratio trans- oil calculate the effective vertical stress.
lation from the overburden to lateral. However, tectonic phenomena may result 2. Calculate the minimum horizontal stress at 3000 m depth for the formation described
in additional horizontal stress components and, thus, two horizontal stresses, one in Part 1. Assume that the Poisson ratio is 0.25. Will the absolute minimum stress
increase or decrease during reservoir depletion?
minimum and one maximum, can be considered, such that
3. Assuming that the Poisson relationship is in effect and assuming that the. horizont~1
stresses are "locked" in place, what would the critical depth be above WhICh a hori-
(3.14) zontal fracture would be generated if 500 m of overburden were removed by some
geologic means?
where O"leCI is the net tectonic stress component.

Solution
Example 3.1 Stresses on a Fault and Principal Stress Magnitudes
and Directions 1. The absolute vertical stress is equal to the overburden pressure, and from Eq. 3.6

Assume that the two-dimensional stresses ax. O"y and 'xy on Figure 3.2 are 3.5 x 107 Pa, a; = PtgH = (2650)(9.8)(3000) = 7.8 x 107 Pa.
4.2 x 107 Pa and 7 x 106 Pa, respectively. Calculate the stresses 0" and T on a fault at an
Since no information on the reservoir pressure is given. a minimum reservoir pressure
angle () = 30. Calculate the principal stress magnitudes and directions in this system.
can be assumed to be simply hydrostatic:

Solution
p "'" PogH = (850){9.8)(3000) = 2.5 x 107 (Pa),

and, thus, from Eq. 3.8


The stresses a and T on a fault at an angle B = 30 can be calculated as follows: From
Eq.3.2 a~= 7.8 x 10' - (0.72)(2.5 X 107) ~ 6.0 X 10' Pa.

a = (3.5 x 107) cos2 30 + (2)(7 x lW)sin 30 cos 30 + (4.2 X 107) sin2 30 2. Equation 3.13 provides the absolute minimum horizontal stress as a function of the
vertical stress. The vertical stress is equal to the weight of the overburden. Furthermore,
= 4.28 X 107 Pa. the pore pressure has been assumed to be hydrostatic. Thus, at 3000 m,

From Eq. 3.3


ah.min = (1-0.250.25) [7.8 x 107 _ (0.72)(2.5 X 107)] + (0.72)(2.5 X 10')

= 3.8 X 10' Pa,

To calculate the principal stress magnitude, first the direction of the principal stresses The effective minimum stress is then
vs the direction of a)" can be calculated.
From Eq. 3.4 a'h.mm. =38 ~ X 10' - (0.72)(2.5 X 107) = 2.0 X 107 Pa.

e = ! tan"! ( (2)(7 x 106) ) = -31.7. During the reservoir pressure depletion, Eq. 3.13 can be used to examine the effects on
2 3.5 x 107 - 4.2 X 107 the absolute horizontal stress:
v
i.e. to the left of the y axis in Figure 3.2. ah = --(ov
1-v
- ctp) + etp.
~

Near-wei/bore stresses 59
58 Stresses in formations

If the reservoir pressure changes CJ.P, then the horizontal stress will change by The problem can be solved also by the simultaneoussolution of

CJ.O"h.min = (a - ~)1-v tip, O"h.m,n = G:~~) [(2650)(9.8) - (0.72)(850)(9.8)]H + (0.72)(850)(9.8)H

and for this problem, = 1.27 x 104H Pa.


and the new vertical stress equation
(0.72)(0.25)
CJ.O"h.min = ( 0.72 - 0
1- .25
tip = O.48tip Pa. a; = (2650)(9.8)(H - 500) = 2.6 X 104(H - 500) Pa.

From the expression above it is obvious that as the pore pressure decreases, the Setting O"v = O"h.mie, critical depth H = 977 m from the "original" ground surface or
decreases. For a 5 x 106 Pa reduction in the reservoir pressure, the minimum
O"h.min 477 m from the current. 0
horizontal stress for this formationwill decreaseby

tiO"h,min:; (0.48)(5.0 X 106) = 2.4 X 106 Pa. 3.3 Near-wellbore Stresses


The magnitude and orientation of the in situ stress field can be altered locally,
3, Figure 3.3 provides a graphical solution to the problem. While the "original"
minimum horizontal stresses remain largely the same, the weight of the overburden as a result of the drilling of a well. These induced stresses often result in stress
has been reduced and the vertical stress profile vs depth has been shifted to the left. concentrations that are significantly different from the original values.
The slope remains the same since the density is the same. The critical depth from this The drilling of a borehole distorts the preexisting stress field. Figure 3.4 offers a
construction happens at the intersection of the original horizontal stress and the new description and conventional nomenclature for the far-field stresses (Figure 3.4a) and
vertical stress.This is at approximately1000 m from the originalsurfaceor 500 m from the stresses in cylindrical coordinates created by the drilling of the well (Figure 3.4b).
the current. At depths shallowerthan this a hydraulicfracture is likely to be horizontal; In all cases the first subscript denotes the direction of the force and the second denotes
at larger depths the fracture will be vertical. the plane upon which the force acts. The index of the plane corresponds to the normal
Interesting phenomena can happen around this critical depth, where a fracture may direction. Thus, stress 0"16 means a stress along the r direction and normal to the 9
initiate horizontallyor vertically and then may tum by 90. This would be caused by plane. In all cases normal stresses with two identical subscripts can be replaced by
an additional fracturepropagation-inducedpressure. a single subscript.
If the rock is assumed to remain linearly elastic and the borehole is drilled parallel
0 to one of the principal stress directions (e.g. a vertical well) the following expression
can be obtained for the near-well stresses. (Note: In this section stresses are effective.)

-500
E 0
a, = 4(0',. + O"y) (1 - ;) + 4(0',. - cry) (1 - 4;; 3~!
) cos(29), (3.15)
8
~
-1000

(1 + ~) -
;)
-500
...
VI

c: O"e = !(O"x + O"y) !(O"x - O"y) (1 + 3;) cos(28) (3.16)


'"e
""
OJ
c:
-1500 -1000
E
'0 S and
'I:
0
E -2000 -1500
0.
."
0 rr()
]
= -2(O'X - O"y)
(2r;
1 + 72 - 3r!)
-;=4 . (28) SIll . (3.17)
g
~ At the borehole wall, where r -+ rw, Eqs. 3.15 to 3.17 simplify to
~ -2500 -2000

a, = 0, (3.18)
-3000
0 20x10" 4Ox10" 60x10" 8Ox10"
(1e = (O"x + (1y) - 2{O"x - O"y)cos(2e), (3.19)

Stress, Pa
and
Figure 3.3 Criticaldepthfor horizontalhydraulicfracturesfor Example3.2 (3.20)
60 Stresses in formations
Near-weI/bore stresses 61

z Considering only the directions parallel and perpendicular to the minimum horizontal
e e
stress directions, i.e. = 0 and = 1</2, Eq. 3.19 simplifies further:

(3.21)

and
y (3.22)
x
Example3.3 Calculation of the Stressesat the Borehole

A well has been drilled in a reservoir where the pore pressure is 2.0 x 107 Pa and the
absolute minimum and maximum horizontal stresses are 2.4 x 107 Pa and 3.6 x 107 Pa,
respectively. Calculate the effective stresses ae at = 0 and e e=
:rc/2. Assume that
ct=1.

Solution
(a)
First, from Eq. 3.8 the effective far-field stresses are

ax = 2.4 x 107 - 2.Q X 107 = 0.4 X 107 Pa,


and
O"zz
Z ay = 3.6 x 107 - 2.0 X 107 = 1.6 X 107 Pa.

From Eqs. 3.21 and 3.22 the effective stresses at e = 0 and e = rr/2 are

r,,~ rzr and

Suppose that in Example 3.3 the tensile strength of the rock was approximately
equal to 4 x 106 Pa. The calculation shows the possibility for tensile failure to occur
in a direction perpendicular to the minimum principal stress solely as the result of
drilling the borehole. Such a value of tensile stress is in the range of typical tensile
strengths of reservoir rocks.
It should be noted that these induced stresses diminish rapidly to zero, away
from the wellbore as shown in Figure 3.5 (two-dimensional view) and Figure 3.6
(three-dimensional view). While the two-dimensional view of Figure 3.5 provides
r,o the stresses in one direction, Figure 3.6 provides their profiles around the well. As
can be seen 0"1) (Figure 3.6a) and 0", (Figure 3.6b) have several minima which would
(b) provide the "venue" for the second branch of a fracture. Once initiated from a point
or a plane of the well the fracture starts propagating in a first branch. Friction pressure
Figure 3.4 Description and nomenclature for (a) Cartesian and (b) radial stress components resistance and other retardation effects could allow the fracture to find another point
of minimum stress concentration for a second fracture branch. Depending on the
required path length (and thus the perforation phasing) it is conceivable that in
certain cases a second branch may not develop. The near-well stress concentration
._._--_._ ... ----

62 Stresses in formations
Stress concentrations for an arbitrarily oriented well 63

120x10
affects greatly the near-well geometry and contact between the fracture and the well.
".= 3.3x107 Pa
"y = 3.3X107Pa
Away from the well the far-field stresses take over.
100x10
Uz = 3.3X107 Pa
'w= 0.1 m

K
' aOx10' o
'"
c.. 3.4 Stress Concentrations for an Arbitrarily
.,
rti
60x10' Oriented Well
~
ij)
(fe
40x10 The solution for the stresses and displacements because of an infinitely long circular
", hole in a homogeneous, isotropic, linearly elastic medium is given by the superpo-
20x10
~( sition of Kirsch's solution, the antiplane solution, and the solution for an internally
pressured hole (Figure 3.7, Deily and Owens [4], Bradley [5], Richardson [6]).
0 To represent this solution mathematically, it is necessary to define the orientation
0.1 10 100 1000
of the borehole with respect to the in situ stresses. Define a as the angle between
Radial distance from welJ, m
the ax(ah,min) direction and the projection of the borehole axis onto the ax - Oy
Figure 3.5 Stress concentration around and away from a wellbore plane, and f3 as the angle between the borehole axis and the 0::. direction as shown
in Figure 3.8. The rotation of the stresses from the in situ system of coordinates to
the borehole local system of coordinates is obtained from the following:
(a) a;a: sin2 f3 cos2 f3 cos2 a cos2 f3 sin2 a
ayy 0 sin2a cos2a
Ozz
T~

TXl
=
cos- f3
0
- sin f3 cos f3
sin 2 f3 cos? a
- sin a cos a sin f3
sin f3 cos f3 cos2 a
sin 2 f3 sin2 a
sin a cos a sin f3
sin f3 cos f3 sin 2 a
{::}
ay
(3.23)

Txy 0 - sin rxcos rxcos f3 sin rxcos rxcos f3


where ax, 0Y' and az are far field stresses.

z
)-,

(a) (b) P.... (c)

Figure 3.6 Three-dimensional view of stress concentration around the wellbore Figure 3.7 Problem decomposition: (a) Kirsch's problem, (b) antiplane loading, (c) internally
pressured hole
__
.. _---_.- _-- --------_ ... _._--_._-

64 Stresses in formations Vertical well breakdown pressure 65

At the borehole wall (r = rw), these expressions simplify to

(3.30)
0'(18 = (axx + ayy - p",) - 2(axx - ayy) cos(28) - 4rxy sin(2e), (3.31)

au ~ au - 2v(axx + ayy) cos(28) - 4VTxy sin(28), (3.32)


rrfJ = = 0,
Tn (3.33)
r~ = 2(-rxz sine + TyZ cos 8). (3.34)

The solution above can be further simplified, with little loss of accuracy, by
assuming the Poisson ratio, v::: 0 (Yew and Li [7]). The major difference with a
borehole (horizontal or otherwise) drilled parallel to a principal axis is that one of
the shear components, r(Jz, remains finite at the borehole wall. The magnitude of this
shear component will affect the overall stability of the borehole. More importantly,
the principal stress tensor will be rotated in the neighborhood of the circular opening.
.: :.: " In other words, the stress condition at the borehole wall will differ in magnitude and
in orientation from the far-field conditions.

Figure 3.8 Pertinent parameters for inclined wellbore geometry


3.5 Vertical Well Breakdown Pressure
The stress field resulting from the in situ stress tensor and the internal borehole
pressure Pw is given by the following: For a vertical well, one of the principal stresses, the vertical stress, is parallel to
the borehole axis. If this stress is the minimum stress, the fracture will initiate and

a.; = 11(0'= + ayy ) (1 ~) - r2 1 D


+ l(a - ayy) (42
1- rZw+ 3r4)
rt cos(28)
propagate horizontally. However, if the vertical stress is not the minimum, then the
fracture will propagate vertically and normal to the minimum horizontal stress. The
breakdown pressure for an uncased, smooth borehole can be written as
4r2 3r4) r2
+ Tx)' ( 1- r2w+ r4w sin(28) + r; P, (3.24) breakdown pressure - stress concentration at the borehole
= tensile strength of rock.

al)f)= !(axx + ayy) (1 + :; ) - 4(axx - ayy) (1 3~!)


+ cos(28) or
Pbd = stress concentration at the borehole + tensile strength of rock.
3r4) 2 The upper limit of the stress concentration at the borehole can be calculated by
-rx),
( 1+ r4w Sin(28)-:;Pw, (3.25)
Eq, 3.22 and, therefore,

(3.26) Ph.upper = 3ax - ay - P + T with ay ::::fJ",. (3.35)

Equation 3.35 is well known and often referred to as Terzaghi's criterion [8]. The
rrf}-_ [ -Zaxx-ayy)sm(28)+rxvcos(28)]
1(. ( 1+---2"; 3r!) (3.27) absolute stress is used in Eq. 3.35. In terms of effective stresses, a~and a;,
. r2 r4 '
Pb.upper ::: 3a~ - a~ + P + T, (3.36)
(3.28) where P is the formation pore pressure and T is the formation tensile strength.
This equation is valid only in the case of no fluid penetration (Detournay and
Cheng [9]) and it gives an "upper bound" for the breakdown pressure. Also, it
(3.29) assumes that the initiation and propagation directions are identicaL However, if
66 Stresses in formations
Breakdown pressure for an arbitrarily oriented well 67
leakage occurs prior to breakdown, Eq. 3.36 becomes more complex and it is neces-
sary to define a "lower bound" for the breakdown pressure:

p _3ax-ay-2rJP+T
b,lower _ 2( , (3.37)
1- v)
where Tj is a parameter, defined by

a(l - 2v)
Tj = ~-:--~ (3.38)
2(1 - v) .

It should be noted that an increase in the pore pressure in the vicinity of the
well corresponds to a decrease in the effective stresses of the rock, and hence a
decrease in the breakdown pressure. Therefore the use of low-viscosity fluids and/or
low pumping rate may decrease the pressure for breakdown.

Example 3.4 Breakdown Pressure for a Vertical Well

A reservoir is 3000 m deep, the absolute vertical stress, C7v, is 7.8 X 107 Pa, the absolute
minimum horizontal stress, C7h,min, is 3.8 X 107 Pa, and the absolute maximum horizontal
stress, O"h.ma>, is 5.0 X 107 Pa. The reservoir pressure at this depth is 2.5 x 107 Pa and Figure 3.9 Stresses at the borehole wall of an arbitrarily inclined well
the tensile strength is 4.0 x 106 Pa. Calculate the open hole breakdown pressure.

They are defined with respect to the local (borehole) system of coordinates in the
Solution following manner:

The breakdown pressure for open hole can be calculated with Eq. 3.35, where 0'1is in the radial direction;
is tangent to the borehole wall (Figure 3.9) and deviates from the borehole axis
0'2
O"x = O"h.min = 3.8 X 107 Pa and C7y = O"h.max = 5.0 X 107 Pa. by an angle
Thus,
y = :21tan -1 ( 2rez ) , (3.42)
O'(j() - azz
and
The fracture will initiate and propagate perpendicular to the least resistance, i.e, the 0'3 is also tangent to the borehole wall and deviates from the borehole axis by
minimum horizontal stress. 0 (900 - y).

Therefore, the trace of the fracture at the borehole wall forms an angle y with
3.6 Breakdown Pressure for an Arbitrarily Oriented Well the borehole axis. This has been observed experimentally by Kuriyagawa et at. [11].
It must be emphasized that the tests mentioned by these authors were conducted at
Tensile failure will occur when the minor principal stress (in tension) reaches the a slow pumping rate, which favors multiple fracture development. At higher rates,
tensile strength of the medium. The principal stresses at the wall of the borehole are fracture coalescence will probably occur almost instantaneously, aligning itself with
given by Daneshy [10]. the borehole axis, near the wellbore.
The location of the failure on the borehole wall, e and the breakdown pressure
(3.39) are obtained by solving for Pw using

(3.40) 0'3 = -T+ p, (3.43)

where p is the pore pressure at the considered point, and by minimizing Pw with
(3.41)
respect to e.
-_ ....--- ------_."-----"--"."._---_._,, "'---" __
. ... -----

68 Stresses in formations Umiting case: horizontal well 69

The above calculations allow a specific prediction of the breakdown pressure and !z:.r = -<Tz sin fJ cos fJ + a
y cos fJ sin /3.
the angle of initial fracture growth according to the general principle of energy mini-
!}~ = O.
mization. Such a fracture will ultimately reorient itself to propagate most efficiently,
perpendicular to the minimum principal stress. The stresses in the coordinate system corresponding to the deviated well, <Till}'<T=.,
and !~ are calculated from Eqs. 3.31, 3.32, and 3.34. a3 is calculated from Eq. 3.41.
The calculation of breakdown pressure is tedious and a computer program is used here.
Example 3.5 The Impact of Non-vanishing Shear Stress on For a given well deviation angle /3, Pbd is calculated by minimizing Pw with respect to
Breakdown Pressure e using
a3 =
-T+ p
Use the data in Example 3.4 (but with <Th.min = 4.2 X 107 Pa) and calculate the break-
. down pressure of a deviated well. Assume that the angle fJ changes from O (vertical The results are plotted in Figure 3.10. 0
well) to 90 (horizontal well) but the well stays on the plane normal to the minimum
horizontal stress.
3.7 Limiting Case: Horizontal Well
Solution For the case of a horizontal well, production engineering considerations have identi-
fied two limiting cases [12]. The first possibility is for the well to be drilled along the
The breakdown pressures are calculated with rnimmum horizontal stress equal to minimum horizontal stress, resulting in transverse hydraulic fractures. The second
Eq. 3.43, by minimizing p", with respect to e. The well stays on the plane normal
option is for the well to be drilled along the maximum horizontal stress resulting
=
to the minimum horizontal stress implying that ex 90. For this particular case,
in a longitudinal hydraulic fracture. Depending on the reservoir permeability one or
au, ayy,au, flY' !v and !%in Eq. 3.23 are simplified:
the other configuration may be desirable. Low reservoir permeabilities would require
the transverse configuration whereas higher to moderate reservoir permeability would
indicate the longitudinal configuration [12].
Equations 3.23 and 3.30 to 3.34 can be adjusted for these two limiting cases. For
a horizontal well the angle f3 = 90, and therefore the tensor in Eq. 3.23 will be
greatly simplified. The three stresses (lz, <Tx, and (ly and can be substituted by (Iv,
(lh.min, and (lh.max respectively. For a well along the minimum horizontal stress the
angle a = 0 and along the maximum horizontal stress a = 90.
With these adjustments to Eq. 3.23 and Eqs. 3.30 to 3.34 the breakdown pressure
60x106 can be calculated using Eqs, 3.40 and 3.41. The breakdown pressure will be the
smallest value calculated from the two expressions.

50x106
Example 3.6 Horizontal Well Breakdown Pressure
<U
a.. 40x106 Use the data in Example 3.5 and calculate the breakdown pressure for two limiting
"ii cases of a horizontal well. In Case 1 the well is drilled along the minimum horizontal
~ stress which would result in transverse fractures, and in Case 2 the well is drilled along
the maximum horizontal stress resulting in a longitudinal fracture.
30x106

Solution
20x106
0 20 40 60 80 100 The stress expressions in the coordinate system corresponding to a horizontal well
(fJ = 90) are simplified from Eq. 3.23
13. degrees
Figure 3.10 The impact of the non-vanishing shear stress on the breakdown pressure of a
well inclined along the fracture plane
..---.-- ... ~.~~-----
----------_. _ .. _._ ..__ -.-------------

70 Stresses in formations
Permeability and stress 71

Solution
rOY= -O"h.min sin a cosa + O'h,ma. sin o cos ,
=
For a horizontal well, f3 90, Using the equations listed in the solution of Example 3.5
rzx = r}:X = O. the far-field stresses are transformed to the coordinate system corresponding to the
horizontal well. With a varying from 0' to 900. the breakdown pressures are calculated
by the same procedures as in Example 3,6, They are plotted in Figure 3,11 as functions
The stresses O"fJ(j, O'::z, and r:;jI at the borehole are calculated from Eqs. 3,31, 3.32,
of the angle a. 0
and 3,33. 0'3 is calculated from Eq. 3.41. The breakdown pressures, Phd, for the two
limiting cases (a = 0' and a =
90') are calculated by minimizing p.; with respect to
e using the formula
3.8 Permeability and Stress
0'3 = -T+ p.
As discussed in Chapter 1, the reservoir permeability plays a crucial role in the
The results are Phd = 5.1 X 107 Pa for a = 0' and Phd = 2.7 X 107 Pa for a = 90,0 production performance of a hydraulically fractured welL It is also well understood
that the permeability tensor exhibits considerable anisotropy (Ramey [14]; Warpinski
and Branagan [15]) which may go unnoticed in radial flow into a vertical well,
3.7.1 Arbitrarily OrientedHorizontal Well Muskat [16], in a classic study, has shown that kx and ky permeability anisotropy
results in an average radial permeability, k = -jkJZ;. Obviously, the same value of
In drilling horizontal wells off of platforms or "drilling pads" it is often not possible can be obtained from an infinite combination of k; and ky values,
to select either one of the two limiting cases. This problem may create interesting A hydraulic fracture which follows a preferential direction and, especially, the
complications from the production of arbitrarily oriented fractured horizontal wells emergence of horizontal wells and the requirement to optimize their drilling direction
(Owens et al. [13]). have forced a new awareness on the issue (Ben-Naceur [17]; Deimbacher et al. [18J).
The angle a formed between the minimum horizontal stress directions and the Warpinski [19] has presented field evidence where horizontal permeability anisotropy
horizontal well trajectories varies between O and 90. The following example shows is as much as 100:1. Anisotropy from 2:1 to 3:1 is considered common, especially
. the impact of a on the well breakdown, in naturally fissured media.
Walsh [201 has suggested that fissure permeability is related to the effective stress
normal to the fissure direction. This idea is based on asperity contacts and an assump-
Example 3.7 The Impact of a on the Breakdown Pressure of a tion of exponential distribution of the asperity heights. Walsh's model was simplified
Horizontal Well by Buchsteiner et al. [21] who arrived at the following expression for the fissure
permeability, kf:
Use the data in Example 3.5 and calculate the breakdown. pressure for an arbitrarily
oriented horizontal well, with a changing from 0' to 90. (3.44)

where a" has been labeled as the closure quality and B tbe conductivity/porosity
60x10"
factor.
The closure quality, a" has been postulated to be affected by the type of fissure
50x10" porosity, degree of mineralization, tortuosity and shape of the flow path. For a single
fissure, the factor B is a strong function of the asperity height and, thus, width of
'"
c,
4Ox10
the flow path. For a system of fissures the factor B can incorporate their cumula-
-0
.0
Q_
tive effect.
If the reservoir permeability can be considered as largely the result of fissures,
30x10" Eq, 3.44 yields

20x10"
k, = [B x
In a; ]
O"y - a p
3 (3.45)
0 20 40 60 80 100
a, degrees and

Figure 3.11 The impact of a on the breakdown pressure of a horizontal well ky = [Byln a;
ax-ap
]3 (3.46)
-
..... ....... _--_._. ----

72 Stresses in formations Measurement of stresses 73

Assuming that a; ~ a; and Bx ~ By (this is not always true, as shown by Buchsteiner half-length, XI' and a real fracture half-length, XI, that obeys
et al. [21]) then the maximum permeability is normal to the minimum stress direction
which would result in the least favorable situation for flow into the hydraulic fracture,
XI = 4fk:, (3.48)
which is also normal to the minimum stress. The permeability normal to the fracture
face is likely to be the minimum value. XI V t;
where k; is normal to the hydraulic fracture and k.. along the fracture.

Example 3.8 Permeability and Stress Anisotropy


From'the results of Example 3.8 the apparent half-length, xI'
would be approxi-
mately 0.75 xf' In certain reservoirs this could cause a substantial reduction in the
expected fracture performance. Thus, stress-induced permeability anisotropy must be
Assume that a; == 0';== 1 X 108 Pa and B, == By == 1.5 (for permeability in md). Calcu- considered in fracture design.
late k, and ky if ax == 6.2 x 107 Pa, uy == 4.8 X 107 Pa, a == 0.72 and p == 3.5 X 107 Pa.
To determine the parameters Bx, By, a; and a;,
Economides et al. [22] have intro-
duced a well-testing technique that uses conventional pressure transient interpretation
Solution methodologies.

From Eq. 3.44


3.9 Measurement of Stresses
kx == Ix 10
[ 1.51n 4.8 x 107 - (0.72)(3.5
8 ]3 _ For hydraulic fracturing applications two stress variables are of the greatest interest:
X 107) - 11 md.
the minimum stress at the target and the direction of the principal stresses. The former
Similarly, from Eq. 3.46 is related directly to the breakdown pressure and the fracture propagation pressure.
The latter controls the ultimate fracture azimuth and has major implications on the
k.= [1.51n
)
1x1OS
6.2 x 107- (0.72)(3.5 x 107)
]3_ - 3.4 md,
near-wellbore fracture path.
In Section 3.2 the three far-field stresses were introduced and related. For the
vast majority of petroleum formations the' minimum in situ stress is the minimum
implying a 3:1 permeability anisotropy. 0 horizontal stress, and hence fractures are vertical (see Example 3.4).
The values of the minimum stresses vs depth, especially in overlain and underlain
layers to the target interval, are essential in predicting both the lateral and vertical
3.8.1 Stress-sensitive Permeability fracture geometry. As will be addressed in Chapter 11 the fracture height greatly
depends on the stress contrasts between the target and adjoining layers.
Following Muskat's permeability definition [16], Buchsteiner et al. [21] have Prats [23] has shown that the differential effective horizontal stress is influenced
presented an expression for reservoir permeability in a "stress-sensitive" formation: by depth, temperature, tectonic strains and pore pressure.
V Ea Edci vEdcj
a; )]3/2
I
k=[BxByln( a; )In( (3.47) dah = --
I-v
d(av - a) + --
I-v
dT + -1 -v 2 + -1 -v 2' (3.49)
ay + ap ax -ap
where E is Young's modulus, a is the poroelastic constant, T is the temperature and
This model was matched with field data from two large and well-known fissured Ci and 6j are tectonic strains in the i and j directions. Obviously, the knowledge of
reservoirs, and the parameters Bx, By, and a;, a;
were obtained. The model of all these variables is difficult and very expensive to acquire.
Eq. 3.47 suggests the possibility of permeability reduction in the life of a reservoir as For hydraulic fracturing what is often more interesting is the closure pressure
the reservoir pressure declines and the effective stresses increase. More importantly, defined by Nolte [24] as the minimum fluid pressure to open an already existing
the nature of permeability anisotropy may change, and, depending on the relationship fracture. This closure pressure may be equal to the minimum horizontal stress if a
between B; and By, a; and a;, it may even "flip-flop" its direction at some time in homogeneous and single layer is fractured. In field applications this is often not the
the life of a reservoir [21J. case because the hydraulic fracture may traverse both lateral and vertical hetero-
These ideas have considerable implications in the planning of horizontal well geneities and other strata. The concept of the closure pressure is intended to account
trajectories, hydraulic fracturing and the design of fracture lengths. Ben-Naceur for all these effects.
and Economides [17J have shown that for the production from infinite conduc- Below, an outline of common stress measurement techniques is presented, as
tivity hydraulic fractures there exists a relationship between an apparent fracture applied to hydraulic fracturing.
.. __ _-----

74 Stresses in formations Measurement of stresses 75

3.9.1 Small Interval Fracture Injection Tests possible and, therefore, subtle changes in the pressure record, aided by the more
sensitive pressure derivative, are employed for the detection of Pis
A widely practiced technique is the successive mechanical isolation of relatively There has been considerable debate as to whether the maximum horizontal stress
small intervals (1 to 3 m) and the injection of small volumes (5 x 10-3 to 10-1 m3) can be determined from these tests. In general, while the concept is plausible,
with each injection lasting 1 to 2 minutes. The important element is the ensuing it requires 'independent analysis of the poroelastic behavior of the rock [26,29].
shut-in pressure decline. The rationale of the technique is based on the concepts Thus, the determination of O"h.max from these tests is considered unreliable and is
introduced by Hubbert and Willis [2J and expounded upon in a series of publications rarely done.
by Haimson and Fairhurst [25-28].
While these tests are particularly useful in open hole completions, they have been
applied to cemented and perforated wells. Care must be taken to assure enough 3.9.2 Acoustic Measurements
open perforations (undamaged) with appropriate phasing. Warpinski and Smith [29J
suggest 13 perforations per meter (actually four per foot) and at least a 90 perforation As an acoustic wave propagates through a rock, it causes a rock deformation. This
phasing, although ideal perforations, of course, should be aligned with the expected deformation affects the "slowness" of the component compressional and shear waves
fracture azimuth. The injected fluid is of low viscosity, usually a KCl solution. (Howard and Fast, [31]). The compressional wave slowness, u-, is an indicator of
The main variable extracted from these tests is the instantaneous shut-in pressure, the rock response to a longitudinal stress. Shear wave slowness, Us. measures the
Pis, which is obtained from the pressure decline following the termination of injec- rock response to a transverse stress.
tion. The Pis is the pressure in the hydraulic fracture at the instant of shut-in. In a Modem sonic logging tools with sufficiently spaced transmitting and receiving
large treatment this value could be considerably larger than the closure pressure and points allow the determination of the sonic wave velocity slowness through the
it will decline to the latter value following fluid leakoff into the formation. reservoir rock (Newberry et al. [32]).
However, because of the very small fracture propagation in the injection tests Both the Poisson ratio and Young's modulus can be calculated from the sonic
described here, and because no masking phenomena such as induced stresses or wave slowness by (Jaeger and Cook, [33])
fracture retardation effects are anticipated, it is postulated that

Pis == Urnin (3.50)


(3.52)
For a vertical fracture in a single layer it can be surmised that

Pis::::::Urnin = Uh.min (3.51) and


Figure 3.12 after Warpinski [30J shows a succession of injection tests with clearly
marked instantaneous shut-in pressures. Often, such a clear demarcation is not E = PUs
2 [3U~- 4U;]
2 2 '
(3.53)
Uc - Us

6
Injection #1 Injection #2 Injection #3 where p is the bulk density of the rock.
5.9 Newberry et al. [32] presented the logging technique using Eq. 3.52 in conjunc-
S tion with measurements of rock densities to calculate horizontal stresses vs depth.
~ 5.8 Their calculation is exactly the one suggested by Eq. 3.13 requiring the knowledge
!b. of pore pressure and, especially, an assumption of the poroelastic constant, Ct.
~ 5.7 This approach has been criticized by several investigators both because of the
'"'"
~ frequent lack of data on poroelastic constants [29] and the use of dynamic acoustic
Q.. 5.6
Pis=5.73x107 Pa measurements in inferring the response of a quasi-static operation such as hydraulic
5.5
fracturing. Yet, while the technique may be flawed in detecting exact stress contrasts
Deplh: 2456.7
102457.3 m between overlain, underlain and target layers, it can clearly delineate deflections in
5.42 the stress values. "Calibration" of the acoustic measurements with independently
3 4 5 6 2 3 4 0 2 3
obtained values from, e.g., an injection stress test, can provide relatively reliable
Time (min) data in specific formations. Logging data are considerably less expensive than injec-
Figure 3.12 Determination of Pis from stress tests (After Warpinski [30]) tion tests.
_- ..._-----_
.... .. _-._-_., ..._------._---- -_ _._.
.. _----
__ ...

76 Stresses in formations Measurementof stresses rr


3.9.3 Determination of the Closure Pressure 103

The determination of the closure pressure is important not only because of its intrinsic
v~lue b~t also because of design considerations (injection pressure demand) and the 'iil
Q.
pressure
,.......-- derivative ~r- .
diagnosis of the net fracturing pressure during the treatment execution. The net cO 10~ .....~ V~ I)
:> Closure
fracturing pressure is defined as simply the fracturing pressure minus the closure 1ij
pressure. Nolte [24] and Smith [34] have suggested relatively qualitative techniques,
:>
c
Q)
.
tempered by field experiences, to determine the closure pressure. The rationale of the
measurement is based on the observation of pressure profiles both during injection
"'0
"'0
c
ell
.
10
and shut-in. The idea is that the pressure patterns are likely to be different while the ~
fracture is open compared to when the fracture is closed.
A c~mmon injection test is the "step-rate" test involving the injection of discretely
escalating rates mto an already initiated fracture. Recommended injection rates start
0.1 10 lQ2
from 1 barrel/min (2.6 x 10-3 m3/s), in increments of 1 barrel/min for about ten
steps. Injection time is equal for each step (1 to 2 minutes) except for the last Time, min
step which may last 5 to 10 minutes. The observed bottomhole injection pressure is
plotted against the injection rate as shown in Figure 3.13. The break point in the plot Figure 3.14 Log-log diagnostic plot of pressure and pressure derivative of a fracture injection
pressure decline (Mayerhofer et al. [35])
corresponds to the fracture extension pressure which is the maximum possible value
for the closure pressure. This extension pressure is usually 7 x lOS to 1.5 X 106 Pa
larger than the closure pressure because of near-well friction and other phenomena. permeability and the filtercake resistance. They will be described extensively in
The step-rate test is frequently followed by a "flow-back" test, the two tests almost Chapter 8.
:u;,ay_s conducte~ in tandem. The well is flowed back immediately after the step-rate Frequently, in the Mayerhofer et al. [35] log-log diagnostic plot, the closure pres-
lllJe~tlOn.To aVOIdrate transient effects masking the pressure response, the flow-back sure can be detected as a sharp change in the pressure derivative curve as shown in
test IS conducted at a rate about 1/6 to 1/4 of the last injection rate. The bottomhole Figure 3.14. The explanation is simple, Before closure, an open fracture provides a
pressure falloff is observed and plotted against time as shown in Figure 3.13. The considerably different pressure response than that from the reservoir which would
pressure decline will show a characteristic reversal of the slope, with the inflection be controlling once the fracture is closed. Pressure derivatives are far more sensitive
point coinciding with the closure pressure. The test is repeated several times to ensure than the corresponding pressure in detecting such phenomena.
reproducibility of the results.
Recently, Mayerhofer et al. [35] have introduced a robust model for the pressure
transient analysis of fracture injection tests. The main purpose of these tests is to
3.9.4 Core Stress Measurements
estimate the controlling variables for fracturing fluid leakoff, namely the reservoir Measurements of stresses using native cores require the directional extraction of
the cores and extensive laboratory equipment and instrumentation. Techniques fall
under two general categories, acoustic, such as the Differential Wave Velocity Anal-
ysis (DWVA) and, static, such as the Anelastic Strain Recovery (ASR) and the
Differential Strain Curve Analysis (DSCA). The latter two offer the possibility to
Fracture extension measure the relative (and, at times, the absolute) magnitudes of the principal in situ
pressure Injection pressure stresses. This section presents a brief synopsis of the rationale and the methodology
~ of interpretation of ASR and DSCA. .
~""" P""'" Friedman [36] has suggested a mechanism for the origin of the in situ state of
~ stresses which is the essence of the core analysis methods. As grains are buried and
undergo lithification, they are compressed and distorted. They are, thus, stressed.
Cementing material percolates through the porous medium, sets and holds the stressed
Injection rate Time grains in their distorted shapes. This situation leads to large amounts of stored
Figure 3.13 Closure pressure determination from step-rate and flow-back tests (After Nolte [24J energy, which is likely to vary depending on the amount of stress exercised in
and Smith [34]) . each direction.
--- ----- ._--_. --------._--_.. _-_ .. ---------_ ...... _._------ ._----------

78 Stresses in formations Measurementof stresses 79

Grains may undergo changes in the state of stresses in geologic time. While the DSCA is essentially a reverse procedure from ASR by measuring the induced
current state of stresses is controlling, historical evidence, such as the presence of relative strains under a confining pressure (Ren and Roegiers, [40]). Measurement
certain fissures, may reflect past states of stress. of strains and calculation of the principal strains similarly to Eqs. 3.54 and 3.55 lead
Thus, strain relaxation, related to the release of stored energy, is presumed to be to the following solution for the three principal stresses, 0'1, 0'2 and 0'3
proportional to the present state of stress. Oriented relative strain relaxation allows (3.61)
the determination of the oriented stresses' relative magnitudes. The maximum strain 0'1 - a p = Clle1 + C12E2 + C13e3,
direction coincides with the maximum stress direction. 0'2 - a p = C12e1 + C1lS2 + C13e3. (3.62)
Supposing that three strain gauges on a horizontal plane, at 0, 45 and 90, (3.63)
measure strains eo, S45 and c9O, then the principal strain directions are (Obert and 0'3 - a p = Cl3el + C13c2 + C33e3.
Duvall [37])
where

(3.54) en = . (3.64)
(1+ v1z)(1 - V12 - 2V31V13)
and E 1(V!2 - V31V13)

ez = So + S90 -
2
.../2/ (co -
-2 S45)2 + (S45 - E90)2
.
(3.55)
C12 =
(1 + V12)(1 - v12 - 2V31V13)
(3.65)

E3V13 (3.66)
The direction of the principal strains is given by Cn = ,
(1- v12 - 2V31V13)

e = 2:1 tan -1 {245 - EO - E90 }


. (3.56) E3(1 - V12)
(3.67)
So - S90

Blanton [38] has solved the problem of the stress magnitudes from the principal
strains and the value of the vertical, overburden, stress O'v:
3.9.5 Critique and Applicability of Techniques
0'1= (O'\)-ap )
(1 - V)LlSl + vCl\S2 + Llsv) +ap (3.57)
(1 - v)l\sv + V(LlS2 + LlSl) Section 3.9 presented the most established techniques for the measurement of stresses
and stress directions that have an impact on hydraulic fracturing. Other techniques
(1 - V)LlS2 + V(l\Sl + l\sv)
0'2= (av-ap ) -s a p, (3.58) exist and there is reasoning in outlining the present methods.
. (1- v)l\sv + v(l\sz + l\sJ) The determination of closure pressure with an injection test is by far the most
In Eqs. 3.57 and 3.58, Sv is the overburden strain. reliable technique, and the measurement is important for the analysis of the fracture
A viscoelastic model, using time-dependent analysis, was constructed by Warpin- net pressure, i.e. the proper determination of the propagating fracture ge?metry.
ski and Teufel [39] for the radial and vertical strains, respectively, er(r) and Sv(/), Injection stress measurements along a vertical column are important to diagnose
using strain values from any angle, e. The forms of their solution are intralayer stress contrast and, thus, predict fracture height migration. ~onic logs,
properly calibrated, can be a reasonable and relatively inexpens~ve substItute.. .
srCt) = (20'1 cos2 e + 20'2 sin2 e - 0'1 sin2 e- 0'2 cos2 e- av)h (1 - e-t/t1) The determination of stress directions in the target interval IS for the prediction
+ (0'1 + C12 + av - 3p)]ZC1- e-t/tl) (3.59) of the fracture azimuth. This is essential information in the hydraulic fracturing of
highly deviated and horizontal wells. ASR and DSCA can readily fulfill this t~k.
and The various tests done together for critical wells can provide all stress magnitudes
sv(t) = (2av + (0'1 + 0'2 + o; - and stress directions.
- 0'1 - (2)J 1(1 - e-t/tt) 3 p )]z(1 - e-t/tI), (3.60)
Clearly, not all tests are necessary in every well that is to be hydraulically =:
where J 1 is the distortional creep compliance, J 2 is the dilatational creep compliance tured. In geologies that are not chaotic, a set of cores from one well may be sufficient
and tl and t2 are the relaxational time constants for the distortional and dilata- for the entire field. Also, in. reservoirs where continuity of layers is established from
tional creeps, respectively. These are rock properties that can be determined in the logs in interwell correlation, stress tests can be done in only a few select~d wells.
laboratory. .Fracture closure tests should be done in almost all wells to be hydrauhcally frac-
From Eqs. 3.59 and 3.60, measurement of strains at different angles e, and for tured. The closure pressure could vary considerably. Conveniently, these tests are
different times, would lead to the calculation of 0'1. 0'2 and a., often part of an overall injection strategy for leakoff determination.
--- -----------_.. -,,_,,---- _---_ _ _--- ----- --- .

80 Stresses in formations References 81

23. Prats, M ..: Effect of Burial History on the Subsurface Horizontal Stresses of Formations
References
Having Different Material Properties, SPEJ, (Dec.), 658-662, 1981. ..
24. Nolte, KG ..: Fracture Design Considerations Based on Pressure AnalYSIS,SPEPE, (Feb.),
1.. Roegiers, J..e..: Elements of Rock Mechanics, in ReservoirStimulation,M.J. Economides 23-30, 1988.
and K.G. Nolte (eds ..), Prentice Hall, Englewood Clift, N.J., 1989. 25. Haimson, B.e. and Fairhurst, C.: Initiation and Extension of .Hydraulic Fractures in
2. Hubbert, M.K. and Willis, D.G.: Mechanics of Hydraulic Fracturing, Trans.AlME, 210, Rocks, SPEJ, (Sept.), 310-318, 1967. .
153-166, 1957. 26. Haimson, B.C. and Fairhurst, c.: Hydraulic Fracturing in Porous Permeable Matenals,
3.. Biot, M.A: General Solution of the Equations of Elasticity and Consideration for a IPT, (July), 811-817, 1969. .
Porous Material, 1. Appl. Mech.., 23, 91-96, 1956. 27. Haimson, B.C.: The Hydrofracturing Stress Measuring Method and Recent Field Results,
4. Deily, F.H. and Owens, T.e.: Stress Around a Wellbore, Paper SPE 2557, 1969.. Intl.J, Rock Mech. Min. Sci., 25,167-178,1978.
5. Bradley, W.B.: Failure of Inclined Borehole, 1. Energy Res..Tech., 233-239, 1972. 28 .. Haimson, B..C.: Confirmation of Hydrofracturing Results Through Comparisons with
6. Richardson, R.M.: Hydraulic Fracture in Arbitrarily Oriented Boreholes: an Analyt- Other Stress Measurements, Proc. 22nd US. Rock Mechanics Symposium Massachu-
ical Solution, Proc. Workshopon HydraulicFracturingStressMeasurements, Monterey, setts Inst.. of Technology, Boston, (June), 379-385, 1981. .
California, (Dec.), 1981. 29. Warpinski, N.R. and Smith, M.B.: Rock Mechanics and Fracture Geometry, III Recent
7. Yew, c.a, Li, Y.: Fracturing of A Deviated Well, SPEPE, (NOV), 429-437, 1988. Advances in Hydraulic Fracturing, Gidley, J.L. et at. (eds.), SPE Monograph 12, SPE,
8.. Terzaghi, K: Die Berechnung der Durchlassigkeitsziffer des Tones aus dem Verlauf der Richardson, TX, 1989..
hydrodynamischen Spannungserscheinungen, Sber, Akad. Wiss., Wien, 132, 105, 1923. 30. Warpinski, N.R.: In-Situ Stress Measurements at U.S. DOE's Multiwell Experiment Site,
9. Detournay, E. and Cheng, AH-D.: Poroelastic Response of a Borehole in a Non- Mesaverde Group, Rifle, Colorado, (March),IPT., 527-537, 1985. .
hydrostatic Stress Field, Int. 1. Rock Mech., Min. Sci. and Geomech. Abstr., 25 (3), 31. Howard, G.C. and Fast, C.R.: Hydraulic Fracturing, SPE Monograph 2, SPE, Richard-
171-182, 1988. son, TX, 1970.
10. Daneshy, AA.: Experimental Investigation of Hydraulic Fracturing Through Perfora- 32. Newberry, B.M., Nelson, R.F. and Ahmed, U.: Prediction of Vertical Hydraulic Fracture
tions,IPT, (Oct.), 1201-1206, 1973. Migration Using Compressional and Shear Wave Slowness, SPEIDOE 13895, 1985..
11.. Kuriyagawa, M., Kobayashi, H., Matsunaga, I., Yamaguchi, T. and Hibiya, K: 33. Jaeger, J.C. and Cook, N.G.W.: Fundamentalsof Rock Mechanics, Chapman and Hall,
Application of Hydraulic Fracturing to three-dimensional In-Situ Stress Measurements, New York, 1979.
Proc .. Second Int. Workshopon Hydraulic FracturingStress Measurements, HFSM 88, 34. Smith, M.B.: Stimulation Design for Short, Precise Hydraulic Fractures, SPEJ., (June)
Minnesota University, (June), 307-340, 1988. 371-379,1985.
12. Brown, J.E. and Economides, M. J.: An Analysis of Hydraulically Fractured Horizontal 35. Mayerhofer, M..J .., Ehlig-Economides, C.A.. and Economides, MJ.: Pressure Transient
Wells, Paper SPE 24322, 1992. Analysis of Fracture Calibration Tests, Paper SPE 26527, 1993.
13.. Owens, K.A., Andersen, SA. and Economides, M.J.: Fracturing Pressure for Horizontal 36. Friedman, M.: Residual Elastic Strains in Rocks, Tectonophysics,15,297-330,1972.
Wells, SPE 24822, 1992. 37. Obert, L. and Duvall, W.E.: Rock Mechanics and the Design of Structures in Rock, John
14. Ramey, H.J. Jr..: Interference Analysis for Anisotropic Formations - A Case History, Wiley New York, 1967. .
IPT, (Oct), 1290-1298, 1975. 38. Blanton, T.L.: The Relation Between Recovery Deformation and In-Situ Stress Magru-
15. Warpinski, N.R. and Branagan, P.T..: Altered Stress Fracturing, JPT, (Sept.), 990-997, tudes, Paper SPE/DOE 11624, 1983. . . .
1989. 39. Warpinski, N..R ..and Teufel, L.W.: A Viscoelastic Model for Determining In-SHu Stress
16. Muskat, M.: The Flow of HomogeneousFluids ThroughPorous Media, McGraw-Hill, Magnitudes from Anelastic Strain Recovery of Core, SPEPE, (Au..g.), 273-280, 1989 ..
New York, 1937. 40. Ren, N.-K. and Roegiers, J.-c.: Differential Strain Curve AnalYSIS- A New Method
17. Ben-Naceur, K.. and Economides, MJ.: Production from Naturally Fissured Reservoirs for Determining the Pre-Existing In-Situ Stress from Rock Core Measurements, Proc.
Intercepted by a Vertical Hydraulic Fracture, SPEFE, (Dec.), 550-558, 1989. Fifth CongressInt. Soc. Rock Mech., Melbourne, 1983.
18. Deimbacher, F.X., Economides, M.J.., Heinemann, Z.E. and Brown, J.E.: Comparison
of Methane Production from Coalbeds Using Vertical and Horizontal Wells, Paper SPE
21280, 1990.
19. Warpinski, N.R.: Hydraulic Fracturing in TIght, Fissured Media, IPT, (Feb ..), 146-152,
1991.
.20. Walsh, J.B.: Effect of Pore Pressure and Confining Pressure on Fracture Permeability,
Int. 1. of Rock Mech., Min. Sci. & Geoph..Abstr., 18, 429-435, 1981.
21. Buchsteiner, H.., Warpinski, N.R. and Economides, MJ.: Stress-Induced Permeability
Reduction in Fissured Reservoirs, Paper SPE 26513, 1993.
22. Economides, M..J., Buchsteiner, H. and Warpinski, N.R.: Step-Pressure Test for Stress-
Sensitive Permeability Determination, Paper SPE 27380, 1994.
4
FRACTURE GEOMETRY

The physics described in Chapter 2 and the analysis of stresses in Chapter 3 dictate
certain conceptualizations of hydraulic fracture geometry. Such conceptualizations
are necessary not only in the modeling of the fracturing process itself but, equally
important, in the prediction of the flow performance of the created fracture.
Simplified geometry is often tractable mathematically. Some of the best known
depictions have been alluded to already in Chapter 2 and will be expounded upon
in this chapter. Deviations from the idealized geometry have given rise to numerical
simulation schemes which purport to account for out-of-plane fracturing, complicated
geometries, multi-layered formations and formation heterogeneities.
There is considerable value in using clearly defined assumptions, consistent laws
and constitutional equations to delineate the basic behavior of hydraulic fractures.
Traditionally, the Griffith-Sneddon (see Griffith [1]; Sneddon [2]) crack has formed
the basis of most fracture geometry models while the work of Hubbert and Willis [3]
has cleared the notion of fracture orientation. The former has been presented in
Chapter 2 while the latter has been dealt with extensively in Chapter 3.

4.1 The Perkins and Kern and Khristianovich and


Zheltov Geometries
Two models, both assuming constant height and two-dimensional (2D) propaga-
tion, have dominated the routine prediction of hydraulic fractures. These are the
Perkins and Kern (PK) [4J and the Khristianovich and Zheltov (KZ) [5] models. In
Section 2.3.4, plane strain was introduced as a 2D (x, y) condition with no displace-
ment in the z direction.
Implicit to the two 2D models is the presumption of plane strain, on the vertical
plane for the Perkins and Kern [4] case and the horizontal plane for the Khris-
tianovich and Zheltov [5J case. Figure 4.1 is a depiction of the two plane stress
conditions. Vertical plane strain, along a fracture, with considerably larger length than
height, allows vertical parallel planes to "slide" against each other. This condition [4J,
further developed by Nordgren [6], is often referred to as the PKN geometry.
---------------
--------------------------

84 Fracture geometry
The Perkins and Kern and Khristianovich and Zheltov geometries 85

The assumption of horizontal plane strain, i.e., an infinite number of "sliding"


Vertical plane
strain condition
parallel planes traversing the height of a fracture, is a plausible simplification for
a short but considerably taller fracture. The horizontal plane strain condition [5]
further developed by Geertsma and deKlerk [7] is often referred to as the KGD
geometry.
The PKN geometry depicted in Figure 4.2 is of an approximate elliptical shape
in both the vertical and horizontal axes. In both directions the width is much smaller
than the other dimensions (of the order of a few millimeters compared to tens or
hundreds of meters.) The elliptical geometry is not entirely true and a more rigorous
treatment will be presented in Chapter 9. The height, hf' is constant and the length,
x f' is considerably larger.
One of the most important characteristics of the fracture is the average width,
W, defined as the total fracture volume divided by the area of one face of the two
fracture wings (or, equivalently, the volume of one wing divided by the area of one
face of the wing.) The maximum width at the wellbore (Section 2.4.2) Wo would
then be approximately related to the average width by

(4.1)

assuming an elliptical shape in both directions. An alternative way of writing Eq. 4.1
Figure 4.1 Vertical and horizontal 2D plane strain condition is to replace (rr/4)2 with a shape factor, y, characterizing the assumed relation-
ship between average and maximum width. This has been used often as 3rr/16.
In Chapter 9 this shape factor will be found to be equal to n:/5 using the original
assumptions of Perkins and Kern [4].
The horizontal plane strain condition of the KGD geometry would result in a
fracture with a rectangular profile at the well as shown in Figure 4.3. Again, the
fracture height, hf is constant. The rectangular shape of a cross section further from
the well has a smaller width, decreasing to zero at the fracture length, x f. The
average width is then related to the maximum width, Wo, simply by approximating
the horizontal cross section with an ellipse:

(4.2)

The PKN and KGD geometries have been used in a great number of engineering
design schemes as reasonable (macro-) approximations of induced fractures. The
hydraulic shape, maintained with the addition of propping materials (propped frac-
tures) or etched by acids (acid fractures) corresponds well with the production-type
models of Gringarten and Ramey [8} and Cinco-Ley and Samaniego [9], described
in Chapter 1.
The same fracture models, along with their limiting approximation (radial) have
been used for the interpretation of observed fracturing pressures and the extraction
of propagation and leakoff variables [10-12].
It should be emphasized here that the two models cannot be used interchangeably.
Figure 4.2 The PKN [4,6] geometry They are mutually exclusive, albeit elegant, approximations. Fractures in formations
The Perkins and Kern and Khristianov/ch and Zheltov geometries 87
86 Fracture geometry

WeT) = _3__pnOVR2 - ,2, (4.3)


nE'
8RpnO (4.4)
Wo = nE' '
16R3 pnO (4.5)
V:;; 3E'
_ 16RpnO (4.6)
W:::':~'

(4.7)

We take the above exact solution as a basis and compare it to an approximate


solution of the problem. The approximate solution is obtained a~p.lying th~ plane
strain assumption in every vertical plane normal to the plane contammg the CIrcle ~
... --",- .. -- ....
shown in Figure 4.4. In other words we compute a line crack ("Sned~on crack"). In
every vertical plane. The characteristic length c of the line crack vanes depending
on the location, x. From elementary geometry

c(x) = VR2 - x2; (4.8)


Figure 4.3 The KGD [5,7] geometry

that are clearly bounded at the top and bottom by lithologies likely to contain the
fracture height could be approximated with the PKN model. Relatively uncontrolled
fracture height or small fracture treatments could be approximated with the KGD
model. In general, KGD-type fractures are not interesting from a production point z
of view.

4.1.1 The Consequences of the Plane Strain Assumption


Both the PKN and KGD geometries are based on the plane strain assumption. It is
natural to ask what are the consequences of such a simplifying assumption. Will the
width volume energy computed from a plane strain assumption be larger or smaller
than the "true" values? Does the vertical or the horizontal plane strain assumption x
give "more realistic" fracture shape? To answer these questions we refer to the results
described in Chapter 2, where we considered mathematical solutions for both the
circular crack and the line crack problems. Our approach here may seem somewhat
theoretical but the conclusions will be of practical importance.
We consider a circular crack (in other words a radial fracture) of radius R, located
in a vertical plane (the x,z plane) and opened by a constant net pressure PnO. The
shape, width, volume, average width and elastic energy of the crack have been given
Figure 4.4 Plane strain approximation to a circular crack
in Sections 2.7 and 2.8:
.--.--- ..-.~ ---------------------------- ----

88 Fracture geometry
Fracture initiation vs. propagation direction 89
therefore from Eq. 2.37

(4.9)

Compare Eq. 4.9 to Eq. 4.3 and keep in mind that r2 = x2 + Z2 reveals' that the
plane strain approximation results in the same shape as the exact solution but the
width is 11:/2times larger everywhere. The error of the plane strain approximation is
ther.efore 57~. Cl~arly the maximum width, volume and average width from the plane
stram .approxIma~IOnare also It /2 times greater than their exact values, respectively.
Knowing the straightforward relation between volume and elastic energy, it is obvious
that the elastic energy estimated from the plane strain approximation is also It /2 times
greater than the exact value given by Eq. 4.7.
The plane strain approximation overestimates the width because it "feels" the
restrictive effect of the tips only in the z direction while it disregards the similar
restric.tion from the x direction. The exact solution of the circular crack problem
takes into account the effect of the tips correctly in every direction.
Intuitively it is obvious that in general a plane strain approximation will overes-
timate the width. The error becomes less if the characteristic dimension normal to
the plane in which the plane strain assumption is applied exceeds considerably the
?the: characteristic dimension. In particular, the KGD geometry is a good approx- (a) (b) (c)
imanon for a short fracture with large height and the PKN geometry is a good
approximation for an elongated fracture with height significantly less than its length Figure 4.5 Fracture initiation from a vertical well (a), turning normal to the least resistance,
in most cases, the minimum horizontal stress (b) and, finally, once the resistance is
(Barree, [13]). In the light of this result the recently appearing practice of creating overcome, evolving into a two-winged fracture (c)
models which additively sum widths calculated from different plane strain approxi-
mations is not correct.
Figure 4.5 depicts a plausible fracture initiation from a vertical well (a) and a
turning fracture (b) en route to its final direction, normal to the minimum hori-
4.2 Fracture Initiation vs. Propagation Direction zontal stress. The illustrations in Figure 4.5 (a and b) are idealizations. Several
near-well cracks are likely to develop coalescing into one dominant fracture. The
As explained in Chapter 3, induced hydraulic fractures away from the well will near-well effects result in additional friction pressure during the fracturing process
propagate normal to the minimum stress. In the vast majority of petroleum appli- (often lumped as "tortuosity effects") and would almost certainly lead to "choke"
cations, this would result in vertical hydraulic fractures, normal to the minimum effects of varying severity during production.
horizontal stress. While some petroleum formations would be shallow enough that The distribution of the induced fracturing pressure would greatly depend on the
a hydraulic fracture could be horizontal (Section 3.2 and Example 3.2) it would dissipation of this energy against the various resistances. In the presence of one plane
be a rare undertaking. Such reservoirs would not be candidates for hydraulic frac- of initiation (e.g. zero perforation phasing) a single wing of a fracture will be created
turing by virtue of their higher permeabilities. In some cases high reservoir pressures first. A second wing will initiate when the resistance in the first wing, because of
may cause complications (see Example 3.2) but these should be considered as rare tortuosity friction losses and other retarding effects, exceeds the fracture initiation
exceptions. pressure of the second wing. The latter will depend on the adhesion between cement
Of course, the near-well stresses in the radial well geometry are different from and casing or cement and formation and the distance the fluid has to travel between
the far-field, Cartesian stress components (Section 3.3.) Fracture initiation at the the perforation and the point of the second wing initiation. The stress profile around
well. is likely to follow a plane that is different from that of the final fracture prop- a well has been shown in Section 3.3 and the point of the likely second wing can
agation. The plane of fracture initiation is also affected greatly by the perforation be predicted. Excessive resistance ahead of a second wing may result in only one
pattern. wing of a fracture since the fracture propagation pressure demand may be lower than
Fracture initiation vs. propagation direction 91
90 Fracture geometry

the friction pressure demand around the wellbore. To avoid this problem perforation
phasing becomes important and a 120 phasing is deemed as minimum. Current
perforating guns are configured routinely with 30 phasing.
Figure 4.5(c) shows a fully propagating two-winged fracture with the final direc-
tion normal to the minimum horizontal stress.
More tortuous paths are to be expected from deviated and horizontal wells.
Although unfractured deviated wells generally have production advantages over
vertical wells, fractured deviated wells, at best, have no advantages over fractured
vertical wells and, at worst, they pose considerable disadvantages. The reason is the
near-well tortuosity.
Thus, in general, it is not recommended to fracture deviated wells, both because of
the non-vanishing share stress components and the invariably larger fracturing pres-
sures and, more critically, because of the aforementioned near-well tortuosity. This
would lead to a highly undesirable reduction in the fractured well production. There-
fore, wells that must be drilled at an angle from a platform or a drilling pad should
be turned vertical in approaching the target formation if they are to be fractured. The Figure 4.6 Transverse vertical fractures from a horizontal well
technology for this type of drilling is readily available.
Horizontal wells can, and at times should, be hydraulically fractured. The
following subsection will describe fracture geometries induced from horizontal wells.
In general, wells to be fractured, from a production point of view, should follow
one of the three principal stresses: vertical or maximum horizontal stresses for
longitudinal fractures or minimum horizontal stress for transverse fractures.

4.2.1 Fractures in Horizontal Wells

Horizontal wells have emerged since the early 1980s as important additions to
petroleum activities, capturing an ever increasing share of all wells drilled.
Production engineering considerations suggest (Brown and Economides [14]) that
many horizontal wells should be hydraulically fractured. In some cases they should
be drilled to accept a longitudinal fracture while, in some others, they should be
drilled to accept multiple transverse fractures. The delineation of this issue and the
optimum well direction based on reservoir characteristics have been dealt with by
Brown and Economides [14].
The difference between an ideal transverse fracture initiated from a horizontal
well (Figure 4.6) and a longitudinal one initiated from a vertical well relates to the
production characteristics. In a vertical well/vertical fracture configuration the flow of Figure 4.7 Turning fracture in a horizontal well from longitudinal initiation to the transverse
fluids is linear from (or into) the reservoir normal to the fracture face and then linear direction [15]
within the fracture. For a horizontal well/transverse vertical fracture configuration,
while the linear part between fracture and reservoir is maintained, inside the fracture
To obtain a geometry resembling the one in Figure 4.6, t~e en~ry.from the well
near the well, the flow reverts from linear to radial, This effect on fractured well
to the fracture must be minimized. Otherwise a configuratIOn ~lmll~r to the one
performance has been quantified [14], and, in general, it results in a considerable
de icted in Figure 4.7 is likely to occur where, even if the well IS drilled properly,
reduction in the flowrate from each individual transverse fracture compared to a
th: fracture will initiate longitudinally and then turn through a very tortuous path to
fracture in a vertical well. However, the composite fiowrate from multiple treatments
become transverse and normal to the minimum horizontal stress.
could be much larger than the one from the single fracture in a vertical well.
.-~~-.--- ..-- -----.---

92 Fracture geometry Fracture profiles in multi-layered formations 93

This is an undesirable event and Deimbacher et al. (15] have shown that a fracture 4.3 Fracture Profiles in Multi-layered Formations
width reduction is likely to occur when the contact length of the turning fracture and
the well is excessive. A simple approximation of this effect is Fracture height migration and the variables affecting it will be addressed in
Chapter 11. A vertical fracture profile can be visualized as in Figure 4.9 where
WI nD the middle is the target interval with fracture growth in the overlain and underlain
(4.10)
WI =ZL' strata. Different elastic properties are likely to result in different fracture widths at
the various layers.
where WI and WI are the widths of the longitudinal-to-transverse and the ideal trans- Fracture height propagation generally is an undesirable occurrence. Beyond the
verse fracture, respectively, D is the well diameter and L is the length of the contact. wasted fracture growth in potentially impermeable layers, there are real dangers for
Equation 4.10 suggests that if L > 1.5D a detrimental reduction to the fracture width a proppant "screenout" during execution. Figure 4.10 depicts a plausible occurrence
is likely to occur. Thus, transverse fractures should be executed through very short when the target interval (e.g., the bottom layer in Figure 4.10) is separated from
open well sections. This can be accomplished either by abrasive jet cutting or multiple another similar layer (top) by a relatively thin layer of larger modulus.
passes with very short perforating guns. Multiple transverse treatments must be done While the fracture height may migrate through the connecting layer, the width
individually with proper zonal isolation. may be such that fracturing fluid can leak off but propp ant may not. Thus, the
A longitudinal fracture from a horizontal well is relatively easier to execute. In fracturing fluid slurry in the bottom target interval may become dehydrated and
most cases this configuration would not offer production advantages over a fractured
vertical well. This would be particularly true in lower-permeability reservoirs, which
are the normal candidates for hydraulic fracturing. However, in relatively higher
permeability formations a longitudinally fractured horizontal well could be more
attractive than a fractured vertical well (Economides et al. [16]).
At times, in offshore operations, it may not be possible to execute one or the
other of the limiting cases in fracturing horizontal wells. Configurations such as
the one shown in Figure 4.8 are then likely. The production behavior of arbi-
trarily fractured horizontal wells has been presented by Deimbacher et al. (15].
The breakdown pressure for these geometries was explained in Section 3.7.1 and
Example 3.7.

Figure 4.8 Vertical fracture initiated from an arbitrarily oriented horizontal well at an angle [

from the minimum horizontal stress Figure 4.9 Vertical fracture profile through a three-layer formation
94 Fracturegeometry
References 95

Figure 4.11 A T-shape fracture

References

Figure 4.10 Fracture height migration and associated width reduction through an adjoining layer 1. Griffith, A.A: The Phenomenon of Rupture and Flow in Solids, Phil. Trans. Roy. Soc.,
of large modulus A 221, 163-198, 1920.
2. Sneddon, I.N.: The Distribution of Stress in the Neighborhood of a Crack in an Elastic
Solid, Proc. Roy. Soc., A 187, 229-238, 1946.
unable to transport the proppant. This would lead to a near-well screenout, rapid 3. Hubbert, M.K. and Willis, D.G.: Mechanics of Hydraulic Fracturing, Trans., AIME,
pressurization and termination of the treatment. 210, 153-166,1957.
Another, relatively common, undesirable occurrence is the T-shape fracture. 4. Perkins, T.K. and Kern, L.R.: Width of Hydraulic Fractures, 1PT, (Sept.), 937-49, 1961;
Trans., AIME, 222, 1961.
Consider a target layer (Figure 4.11) at a depth where the minimum horizontal
5. Khristianovitch, S.A. and Zheltov, YP., Formation of Vertical Fractures by Means of
stress is relatively near the vertical stress. Vertical fracture initiation may result
Highly Viscous Fluids, Proc. World Petroleum Congress, Rome, 2, 579-586, 1955.
in fracturing pressure that may exceed the overburden. If, in addition, the overlain 6. Nordgren, R.P.: A Propagation of a Vertical Hydraulic Fracture, SPEJ, (Aug.), 306-314,
layer is relatively difficult to fracture, a T'-shape fracture may be created in the 1972; Trans., AIME, 253, 1972.
target layer. 7. Geertsma,J. and de KIerk, F.: A Rapid Method of Predicting Width and Extent of
Again, the width of the horizontal branch may be large enough to accept fluid but Hydraulically Induced Fractures, JPT, (Dec.), 1571-81, 1969.
not large enough to accept proppant, leading to a treatment-terminating screenout. 8. Gringarten, A.C. and Ramey, A.J., Jr.: Unsteady State Pressure Distributions Created
Phenomena such as the above have contributed to the need for pseudo-3D and 3D by a Well with a Single-Infinite Conductivity Vertical Fracture, SPEJ, (Aug.), 347-360,
fracture simulation schemes. The elegant but simplified PKN and KGD geometries 1974.
would no longer be adequate. 9. Cinco-Ley, H. and Samaniego, F.: Transient Pressure Analysis for Fractured WelIs,1PT,
(Sept.), 1749-1766, 1981.
96 Fracture geometry

10. Nolte, K.G. and Smith, M.B.: Interpretation of Fracturing Pressures, IPT, (Sept.),
1767-1775,1981.
11. Nolte, K.G.: Fracture Design Considerations Based on Pressure Analysis, SPEPE, (Feb.),
23-30, 1988.
12. Mayerhofer, M.J., Ehlig-Economides, c.A. and Economides, M.J.: Pressure Transient
Analysis of Fracture Calibration Tests, Paper SPE 26527, 1993.
5
13. Bartee, R.D.: A Practical numerical Simulator for Three-dimensional Hydraulic Fracture
Propagation in Heterogeneous Media, Paper SPE 12273, 1983.
14. Brown, J.E. and Economides, M.J.: An Analysis of Hydraulically Fractured Horizontal
RHEOLOGY AND
Wells, Paper SPE 24322, 1992.
15. Deimbacher, F.X., Economides, M.J. and Jensen, O.K.: Generalized Performance of
Hydraulic Fractures with Complex Geometry Intersecting Horizontal Wells, Paper SPE
LAMINAR FLOW
25505, 1993.
16. Economides, M.J. and Deimbacher, F.X., Brand, C.W. and Heinemann, Z.E.: Compre-
hensive Simulation of Horizontal Well Performance, SPEFE, (Dec.), 418-426, 1991.

5.1 Basic Concepts


Matter responds by a finite deformation to external force applied normal to its outer
surface. From this point of view solids and liquids act rather similarly while gases
sustain less resistance, i.e. they are more compressible. This difference of behavior is
more quantitative than qualitative. Though less perceptible, a force acting on a plane
may have a direction parallel to it. The response of a solid to such a force is, again,
a finite deformation. Liquids and gases react to such a force differently, manifesting
a continuous deformation called flow [1-3]. The common name of matter able to
flow is fluid.
It is convenient to consider flow as the sliding of parallel layers relative to one
another. Figure 5.1 illustrates this concept. The external forces originate from the
difference of pressures and/or from gravity (Poiseuille flow) or from torque (Couette
flow). The shear stress keeping the system in equilibrium acts in the opposite direc-
tion and has the magnitude:
F
r= - (5.1)
A
which is force divided by the area.
Since the external force and stress-induced force are balanced, it is sufficient to
use one of them to characterize the state of the fluid. The change in velocity Su is
proportional to the distance of the layers tiy. The limit
. tiu
y=- (5.2)
tiy
is called shear rate. One can consider the shear stress as the response of matter to the
shear rate. For Newtonian fluids the shear stress varies linearly with the shear rate.
The coefficient of proportionality is called viscosity. The larger the viscosity the more
resistant the fluid is to flow. If the linear relation does not hold but the shear stress
is still a unique function of the shear rate, we speak about a general (or generalized
Newtonian) fluid. The shear stress versus shear rate relationship expressed in an
98 Rheology and laminar flow Basic concepts 99

line passing through the origin. If there is a positive shear stress necessary to initiate
flow we call this stress yield stress and the behavior is called plastic. Pseudoplastic
behavior means that the fluid has no yield stress but the slope of the rheological
curve decreases with increasing shear rate. Dilatant behavior means that the slope
increases with shear rate. A real fluid may show a combination of different behaviors
depending on the considered shear rate interval. In addition, the exact behavior at
the two ends of a shear rate spectrum is always subject to transitional uncertainties.
In some cases it may be convenient to use only the rheological curve (or another
graphical presentation of the same information) without a parametrization. In the era
of computer applications, however, it is more straightforward to deal with parametric
Figure 5.1 Sliding layers concept of fluid flow models, i.e. with constitutive equations.
The straight line, passing through the origin in Figure 5.2, corresponds to a Newto-
algebraic form is the rheological constitutive equation and its graphical representation nian fluid with the "constitutive equation"
is the rheological curve. The rheological behavior is a material property, independent
of the geometry of the flow conduit. L = fLY. (5.3)
At increased flow rates, the concept of sliding parallel layers is no longer appli- Fluids whose rheological behavior obeys a constitutive equation other than the
cable because of the appearance of more complex movements leading to turbulence. expression in Eq. 5.3 are referred to as non-Newtonian. Almost all fluids used in
While Newtonian fluids exhibit a marked transition to turbulence, most polymer hydraulic fracturing behave in this manner, except for water.
solutions show a gradual change in the flow regimes. In this chapter we make extensive use of the yield power law model (often referred
to as the Herschel-Bulkley model)
5.1.1 Material Behavior and Constitutive Equations (5.4)

Fluids can be classified by the shape of their rheological curves. Figure 5.2 illustrates where the three constants are ty = yield stress, K = consistency index, and n = flow
some rheological features. A fluid is Newtonian if the rheological curve is a straight behavior index.
Equation 5.4 reduces to the Newtonian model if the yield stress is zero and the
flow behavior index is unity. The limiting case where Ly = 0 leads to an expression in
wide use for the description of polymer solutions which constitute the vast majority
of fracturing fluids. This is the well known power law
(5.5)
Finally, it reduces to the Bingham plastic model, if n is equal to unity. To illustrate
the versatility of the yield power law, some rheological curves generated from it,
are shown in Figure 5.3 (Cartesian coordinate system) and Figure 5.4 (logarithmic
coordinates). Several interesting conclusions can be drawn comparing the two figures.
The linear plot is not a very efficient tool to diagnose the presence of a yield stress.
Conversely, a logarithmic plot deemphasizes the differences at high shear rates. Note
that in the Cartesian system the base case (1) and the power law case (2) seem similar
while on the logarithmic plot the base case (1) and the Bingham plastic case (3) are
almost indistinguishable.
Several other constitutive equations have been suggested in the rheological litera-
ture. It is useful to keep in mind that these idealized models are convenient devices
and that reality may be far more complex. The selection of model should be dictated
by the field of the application, by the accuracy of the measurements and by other
considerations such as computational ease.
Shear rate, y The shear stress/shear rate relationship is not the only way to characterize the
Figure 5.2 fluid types and their rheological curves rheological behavior. In some cases it is more convenient to present the apparent
---- - - _--_._-
.. ---- __ .. _-_.-
. ----- .. -- ._-_ .. --_.

100 Rheology and laminar flow Basic concepts 101

viscosity
r
lAa = 7
Y
(5_6)

130
as a function of the shear rate. Graphically, the apparent viscosity is the slope of a
line drawn through the origin and the point on the rheological curve corresponding
to the given shear rate, as shown on Figure 5.5. Note that the apparent viscosity is
'"
CL
p
60
independent of the geometry of the flow conduit. The same name is used for another
variable defined in conjunction with the Hagen-Poiseuille law (see below). To avoid
oi misinterpretation we will use another name (equivalent Newtonian viscosity) for that
'":!?
w purpose. An apparent viscosity curve can be obtained from the rheological curve by
~Q) 40 applying the definition point by point.
s:
CJ) 1',. (Pal n K(Pas")
Figures 5.6 and 5.7 show the apparent viscosity curves for the four different
1 5 0.8 1.6 cases of the yield power law model considered previously. The typical behavior of
2 0 0.6 1.6
3 5 1 0.2 a polymer liquid is more complex. Figure 5.8 is a logarithmic plot of the apparent
20 4 0 1.8 0.005 viscosity of two different solutions of hydroxypropylguar (HPG) polymer in water.
Note the different low and high shear rate behavior, emphasized by the logarithmic
scale. The straight line in the middle indicates the domain where the power law is
"valid" while at low and high shear rates the horizontal portions show Newtonian-like
o 200 400 600 800 1000 behavior. A possible algebraic representation of the above behavior is the modified
Shear rate, )"(1/S)
Cross model (Chakrabarti et al. [4]);

Figure 5.3 Yield power law model - Cartesian coordinates lAO -lAs + (5.7)
lAo = ( )a lAs,
1 + lAo ~ lAs Y

'r (Pa) n K(Pas")


1 5 0.6 16
2 0 0.6 16
3 5 1 02
4 0 1.8 0005

10'
""'
~

..
(->
I-'
oi .,;
'"~ d>
W
~
.;;
lii
'"
:ll d>
s::.
U5 10' (/)
4 f.J -
a-yT

10-' 10 10' 102 103


Shear rate, Y (1/S) Shear rate. y
Figure 5.4 Yield power law model - logarithmic scale Figure 5.5 Concept of apparent viscosity
-----------------------_ _.-._-_ _-_ ------------~

102 Rheology and laminar flow Basic concepts 103

D.B
4
<i)
m
!!:..,
0.6
~
~
'<i; ., (Pal II K (PaS,,)
0
0 1 5 0.6 1.8
'"
:; 0.4 2
3
0
5
0.6
1
1.8
0.2
E 4 0 1.8 0.005
CD
(ii
C-
o. 3

10-' lCP 10' 10" 103 10'

400 600 BOO 1000


Shear rate. y (1/$)
y
Shear rate, (liS) Figure 5.8 Apparent viscosity of different HPG solutions

Figure 5.6 Apparent viscosity of yield power law fluids - Cartesian coordinates
where J-to= low shear rate viscosity (at room temperature 0.0291 Pa- s for the 0.25
102 ~-------------------------------------------,
mass % HPG solution and 0.410 Pa- s for the 0.5 mass % solution); J-ts = high shear
rate viscosity (0.89 x 10-3 Pa- s), r" = stress parameter (3.72 Pa) and a = exponent
'y (Pal n K(Pas")
1 5 0.6 1.6 (0.641) (Chakrabarti et al. [4]).
2 0 0.6 1.6
3 5 1 0.2
The determination of a rheological curve in graphical form and/or its representa-
4 0 1.8 0.005 tion by an appropriate constitutive equation is the main task of rheometry. While the
10'
<i) shear stress is readily available for measurement, at least at the walls of a device,
m
IJ.. the determination of the shear rate is more complicated. A curve similar to the
rheological curve but constructed from observable quantities is called a flow curve.
Understanding of the relation between a flow curve and the rheological curve is
crucial for any application of rheology including the computations related to pressure
driven flow.

5.1.2 Force Balance


Certain flow conduits obey a specific symmetry leading to a simple velocity profile.
Pressure driven flow between parallel plates and in circular tubes has the impor-
tant property that for a given cross section the wall stress is constant along the
perimeter. The force balance around a' body with cross sectional area, Ac, wetted
let' ,ao 10' 10" perimeter, P, and length, L, provides the basic relation between wall stress and
Shear rate, y(lIS) pressure drop:
Figure 5.7 Apparent viscosity of yield power law fluids - logarithmic scale
LPr = Ac!!"p. (5.8)
------ ---------------_- __ -----. __ .. .__
__ - __ ._--- ----- --------

104 Rheology and laminar flow Slot flow 105

The right-hand side of the equation represents the force driving the flow and the left- For a circular pipe, I is the diameter, and for other geometries it is often called the
hand side is the force arising at the outer surface of the body keeping the balance of hydraulic diameter. The cross section divided by the wetted perimeter is sometimes
the system. If applied to the whole cross section, the equation gives the relation of referred to as the hydraulic radius. (The name is somewhat misleading because for
the wall stress to the pressure gradient a circular pipe it is half of the actual radius.)
While all the other variables are general in the sense that they do not refer to a
.w = (;) ilZ (5.9) specific constitutive equation, the Reynolds number is an exception. It is inherently
connected with Newtonian behavior. Since for non-Newtonian fluids the "viscosity"
If applied to any smaller cross section symmetrically located around the center plane is not unique, several choices are possible to extend the definition of the Reynolds
(parallel plates) or around the center line (circular tube) Eq. 5.8 states that the stress number. Whatever choice is preferred, for laminar flow a friction factor vs Reynolds
varies linearly with the distance from the symmetry axis, since Ac/P is a linear number relation is simply a restatement of the flow rate vs. pressure drop relation
function of that distance and the pressure gradient is a constant. derived from the rheological model. For turbulent flow characterization, however,
Once we know that the stress varies linearly along the coordinate representing the concept of Reynolds number plays a central role.
the distance of a contour line from the symmetry axis, we can obtain the shear rate
and, by integration, the velocity profile. Knowing the velocity profile, a subsequent
integration enables us to determine the flow rate, q. The resulting flow rate vs pressure 5.2 Slot Flow
drop relationship is the basic equation for the given flow geometry. Several additional
5.2.1 Derivation of the Basic Relations
flow characteristics are of interest. These include the maximum velocity, Umax and
the average velocity uavg, where the averaging is carried out along the cross section One way to consider a hydraulically induced fracture is to envision a channel of
available for flow; the ratio of the maximum to the average velocity and the kinetic rectangular cross section with width w and height h, where wlh. _,. O. Figure 5.9
energy correction factor, CtKE. The latter two quantities are different measures for shows the schematic of flow between parallel plates of "infinite" height. The ratio
the flatness of the velocity profile. The kinetic energy correction factor, defined by of the cross section to the wetted perimeter is

(5.10) Ac hw w
-= _,.- (5.14)
P 2(h + w) 2'
will be useful in energy calculations. and the characteristic length to be used in Eq, 5.12 is
The flow rate vs pressure drop relation can be presented as friction factor vs
Reynolds number as well. The Fanning friction factor is defined as the proportionality I =2w. (5.15)
coefficient between the wall stress and a suitable approximation of the kinetic energy In this section we derive flow characteristics for the yield power law fluid and then
per unit volume expressed in terms of the average velocity [2]: apply them for some special cases. The solution strategy is as follows: (1) derive
(5.11) the velocity distribution, (2) calculate the maximum velocity, (3) calculate the flow

In some of the European literature, the Weisbach friction factor is preferred. Denoted
usually by A, the Weisbach friction factor is equal to four times the Fanning friction
factor. In this book we use only the Fanning friction factor and omit the name
Fanning in front of it. The friction factor is dimensionless. The average velocity
means flow rate divided by the cross sectional area, i.e. the averaging is carried out
with respect to the area.
The Reynolds number, N Re, is a dimensionless quantity characterizing the ratio
of inertial and viscous forces:
NRe = pUavgl , (5.12)
f-J-
where the characteristic length, I, is defined by

1 = 4Ac (5.13) Figure 5.9 Schematic of slot flow geometry


P
_._ _--_._---------------------------------::,------

106 Rheology and laminar flow Slot flow 107

rate, (4) obtain the average velocity, and (5) determine the quantities characterizing where u(O) is zero because of the no-slip condition. The solution to Eq. 5.23 is not
the smoothness of the velocity profile. well known in the literature:
For slot flow the force balance takes the form
u(y) = 2(n: 1) {W(l- ) [Tw(lK- )] lin
w6.p
Tw=U' (5.16)

With a linear relationship, the shear stress at a distance y from the wall (and acting _ [w(1- ) - 2y] [T",(1- ~- 2Y/W)fln}. (5.24)
in the direction normal to the flow path) is
The maximum velocity is obtained by substituting y = Yo into Eq. 5.24:

T= (5.17) Umax= 2(n


n
+ 1)w(1 - )
[Tw(1 -
K
)] l/n (5.25)

2 It is instructive to plot the different velocity profiles corresponding to the different


At a certain distance from the wall, Yo, the stress is equal to the yield stress. In cases of the yield power law model studied in Figures 5.3,5.4,5.6 and 5.7. Assuming
the domain inside Yo the stress is not enough to cause velocity change and therefore w = 20 mm and Tw = 25 Pa, the resulting profiles are shown 00 Figure 5.10. The
the fluid moves as a plug with the uniform velocity Urnx. Clearly, constant velocity plug starting at a distance from the wall, Yo = 8 mm, can be well
seen for cases (a) and (c).
Yo = 1- -Ty) -,W (S.18) The total flow rate is the sum of the flow rate in the plug, qi and in the outer
(
Tw 2 part, q2. The plug flow rate is simply
i.e. with increasing ratio of yield stress to wall stress the plug is wider and wider. It qi = hew - 2yo)umax, (5.26)
is convenient to introduce a new variable for this ratio:
while qz is obtained integrating the velocity profile:
Ty
(5.19)
=-,
Tw q2 = 2h 1)'0 u(y)dy. (5.27)
which has to be less than unity; otherwise no flow occurs. In terms of the newly
Substituting Eqs. 5.24 and 5.25 into the latter two equations and carrying out the
introduced variable, Eq. 5.18 takes the form
integration, we arrive at the following expression for the total flow rate:
(1- )w
Yo = (5.20) = hw2n(l- )(1+ n + n) [Tw(1 - )] 11"
2 (S.28)
q 2(I+n)(1+2n) K
The shear rate is simply the derivative of the velocity with respect to the distance
from the wall. Rearranging the specific constitutive equation (Eq. 5.4), we obtain

:; = c~TYr/" (S.21)

Substituting the shear stress from Eq. S.17, we arrive at the differential equation
describing the velocity field:

du _
dy -
[(W/2-
w/2
y
Tw
_
Ty
) /
K
]1/n (5.22)

The velocity profile is obtained by integrating Eq. S.22 from the wall to a given
location y:
Centerline

u(y) - u(O) = Jo
e[(W/2-y' w/2 Tw - Ty
)/ K
]1/n dj/, (S.23) Figure 5.10 Velocity profiles
-------------------------------, """-',,,.,,-,,------_._---------------------- ---~-,,-- ..,,--------------~

108 Rheology and laminar flow Slot flow 109

from which the average velocity is property. Therefore, NRe = 241j, and

( 1_)l/
Uav =!!._ = wn(l- )(1+ n + n) [<w(1 _ )] l/n n
g hw 2(1 + n)(1 + 2n) K
(5.29) 6uavgpw(1 - )n(l + n + n) ~
N~e= (5.35)
Equation 5.29 does not contain the height, indicating that the flow has the same (1 + n)(1 + 2n)<~-1/n
characteristics in any horizontal plane. This is a consequence of the assumption
wf h. -lo- O. The usual measure of the flatness of the velocity profile , umax II'"avg, can From Eqs. 5.4 and 5.19 we can express the shear rate at the wall:
be obtained from Eqs. 5.25 and 5.29,

Umax 1+ 2n
_ _ [(1-
Yw - K
)<w] l/n
(5.36)

uavg
- 1+ n + dm '
(5.30)
Consequently, the apparent viscosity at the wall (wall shear stress divided by wall
where, interestingly, this ratio is a function of nand only. The same is true for shear rate) is
the other measure of flatness, the kinetic energy correction factor
/I = <1-1/n __K ) l/n
(5.37)
.....w w ( 1-
CXKE = _CU_av_g)_3= (2 + 3n )(3 + 4n )(1 + n + n)3
(u3)avg (1 + 2n )2(6 + 18n + IIn + 18n2 + 28n2 + 6n3 + 18n3) The latter two equations enable us to reformulate the expressions for the average
velocity and for the Reynolds number:
(5.31)
wn(l- )(1 + n +n).
where the averaging operation is carried out on the cross sectional area available for (5.38)
flow. uavg = 2(1 + n)(1 + 2n) Yw
The flow rate vs wall stress relationship can be written in several forms. Often I 6uavgpw(l- )n(l + n + n)
(5.39)
the pressure gradient is preferred over the shear stress. Then, a convenient formula- N Re = (1 + n)(1 + 2n)J1-w .
tion is
q= hw2n(l- )(1 + n + n) [W(I- ) f;.P] lin It is obvious that the corresponding relations can be obtained easily for the special
(5.32)
2(1 + n)(l + 2n) 2K L cases of Bingham plastic, power law and Newtonian fluids by substituting the specific
values for the yield power law parameters. Table 5.1 gives a summary of the results.
where When the wall stress (pressure drop) is not known, the corresponding relations
= 2"CyL. (5.33)
shown in Table 5.2 are more convenient to use.
f;.pw In Table 5.2 we also show the relationship between the average velocity and wall
shear rate. The shear rate at the wall for a Newtonian fluid
Unfortunately, we cannot express the pressure gradient as a function of the flow
rate. Si~c~ in en~inee~ng applications we most often wish to calculate the pressure . 6uavg
YwN=-- (5.40)
drop, this is a senous disadvantage, Nevertheless, any suitable root-finding numerical w
method can be applied.
Another form of the same relationship is the combination of Eqs. 5.11 and 5.29: can be deduced from Eqs. 5.36 and 5.3. The right-hand side of Eq. 5.40 is an observ-
able quantity, even if the fluid is non-Newtonian. It is called nominal Newtonian
f = 4(1 + n)(1 + 2n)t~-1/n shear rate. WaH stresses, found experimentally, are often presented as a function of

uavgpw(1 - )n(1 + n + n)
1-
(~
) u:
(5.34) the nominal Newtonian shear rate. Such a curve is called a flow curve.
The general expression of the wall shear rate for the yield power law fluid, in
terms of the nominal Newtonian shear rate, is
The Reynolds number has been defined with respect to Newtonian fluids. The product
j x NRe equals 24 for the Newtonian slot flow (as we will see soon). One possible _ [ 1 + 3n + 2n2 ] 6uavg (5.41)
Yw = --,
way to generalize the Reynolds number for non-Newtonian fluids is to preserve this 3(1-)n(1+n+n) w
Slot "ow 111
110 Rheology and laminar flow

Table 5.1 Bingham plastic, power law and Newtonian formulas for slot flow and therefore the dependence of the wall stress on the nominal Newtonian shear rate
can be given as
+ 3n + 2n 2]
Bingham plastic Powerlaw Newtonian
1 n 6uavg n
T =Kyn r", = r
y
+K
[ 3(1-)n(1+n+</m) (W)--- (5.42)

wn (T.) I/. In the section describing tube flow, we give a detailed discussion of the relation
Um"" = 2(1 +n) K
between the flow curve and the true rheological curve.
= hwzn (TIV)]/. Equation 5.42 is an implicit relationship because is defined in terms of the wall
q 2(1 +2n) K stress. For a Bingham plastic fluid an explicit expression is available for (in terms
wn (Tw)I/n WT..
=-
of the average velocity) as seen from Table 5.2. For a power law fluid, the expression
Uavg = 2(1 + 2n) K Uavg
61l- is explicit because = O.
Urn"" 1+ 2n U",a. 3
-=-
Uavg 2 + t/! U.vg = l+n Uavg 2 5.2.2 Equivalent Newtonian Viscosity
+ )3
35(2 (2 + 3n)(3 + 4n)
aKE = 27(16 + 19t/!) xe = _;_-;6-::(1:-+~2-n::::)2~ Often it is convenient to characterize the fluid by the viscosity of a Newtonian fluid
causing the same pressure drop. This viscosity is called here equivalent Newtonian
hw2n (w i!>p) lIn viscosity (but the terms effective viscosity and apparent viscosity have also been
q = 2(1 + 2n) 2K T used in the literature.) The equivalent Newtonian viscosity (for slot flow) is defined
N' _ 6nwuavgP _ 2wu.vgP by the Hagen - Poiseuille law:
NR<- ---
R.. - (1 + 2n)KI/"-r;-I/n f,L w2 /:!,.p
Me = ----, (5.43)
24 24 12uavg L
/=- /=~ j=-
NR~ - N'p,< ; and it is a property associated both with the fluid and the geometry of the flow
channel. From Table 5.2 it can be expressed both for a Bingham plastic and for a
Table 5.2 Wall shear rate, wall stress and pressure drop for slot flow power law fluid. Table 5.3 shows both the equivalent Newtonian viscosity, J.Le =
rw/YwN and the apparent wall viscosity, J.l.w = rw/yw' Also included is the Reynolds
Bingham plastic Powerlaw Newtonian number. Substituting the explicit expression for from Table 5.2, one can obtain
T=KY' T=Jl.Y explicit expressions for a Bingham plastic fluid. This exercise is left for the reader.

. _ (1~ + 2n) -;-


y",-
6u,vg .
yw=--
6u.vg
W Example 5.1 Flow in Fracture Assuming Slot Geometry

T",=1l- 6uavg) A fracture is h = 30 m high and w =


10 mm wide. Flow of a yield power law fluid
( -- W (ry =
5 Pa, n =
0.6, K = 0.2Pa _5") results in a pressure gradient t:1p/L = 1.8 kPa/m.

6p = 2Ty Table 5.3 Different viscosities of non-Newtonian fluids for slot flow
L w
Bingham plastic Power law
A.
'I' =-
2 'f"'T'b
V 1 + uo cos 4rr + arccos[(1
3
+ bo)-3] xw-(n+1Ju~vg
T=KV

) "-1( )"-1
4f,Lpuavg 1 + 2n 6uavg
( 1_----'-t/!):_:.(2_+_..:..<I>:__)w
where bo = ---
WTy f,L", = Il-p + Ty ..:. 12uavg
/).w=K
( --
3n
--
W

" _ 2"-1
_ (1+2n)"
-- Kwl-"u"-1
f'Ve -3 1l 3'\18

, 2wu.vgP 12u;vgP<P
NRc = --- = ---
Il-e Ty
112 Rheology and laminar flow Flow in circular tube 113

What is the flow rate assuming slot flow? What is the calculated pressure gradient at and, following Eq. 5.13, the characteristic length is D. The force balance applied to
the same flow rate if the yield stress is neglected? (Assume laminar flow. We will see the total cross section gives
in Chapter 6 how to check the validity of that assumption.) Df1p
w= 4[. (5.45)

Solution The other consequence of the force balance is the linear variation of the shear stress.
The shear stress at a distance y from the wall is
From Eg. 5.33

.= (~-y).
2ryL
4> = -- = 0.556;
tipw (5.46)
D Tw
therefore, flow rate is determined (from Eq. 5.32) as 2
Similarly to the flow between parallel plates, at a certain distance from the wall,
= hw1n (1 -.p)(l + n + n4 [weI - 4 c,.p] lin = 0.0324 m3/s (12.2 b m).
Yo, the stress is equal to the yield stress. Within Yothe stress is not enough to cause
q 2(1 + n )(1 + 2n) 2K L P
a velocity change and the fluid moves as a "plug" with the uniform velocity, umax.
At the same flow rate (q = 0.0324 m3/s) but neglecting the yield stress, the pressure The distance Yo depends on the ratio of the yield stress to the wall stress, 4>, and can
gradient from Table 5.2 is be expressed as
(1- )D
Yo = (5.47)
2
The derivation of the relationships follows the pattern introduced for slot flow. The
average velocity is given by
Neglecting the yield stress, but retaining the same K and n, less than one-third of the
originally given pressure gradient is calculated. 0
nD [TwClK- )r/n (1 - )[(2n2 + 3n + 1) + (2n2 + 2n) + (2n2)2]
uavg = 8(1 + n)(1 + 2n)(1 + 3n)
5.3 Flow in Circular Tube (5.48)
The flatness measures take the (somewhat complex) form:
5.3.1 Basic Relations
Umax (1 + 2n)(1 + 3n)
If D is the diameter of the pipe (see Figure 5.11), then the ratio of the cross section ~ = -::--::-----,.-.:.....,..--'-:::-----::------::---;;--;;-
(5.49)
Uavg (2n2 + 3n + 1) + (2n2 + 2n) + 2n2</J2'
to the wetted perimeter is
D and
(5.44)
p = 4'

(5.50)
where

ao = 36n5 + 159n4 + 279n3 + 243n2 + 105n + 18,


al = 108n5 + 390n4 + 522n3 + 306n2 + 66n,

oz = 216n5 + 477n4 + 350n3 + 85n2.

All the equations above are easy to use if the wall stress (pressure gradient)
is known. In engineering practice usually the flow rate (average velocity, nominal
Newtonian shear rate) is specified and the wall stress (pressure gradient) has to
be determined. This requires an iterative solution procedure. Efficient root-finding
Figure 5.11 Schematic of tube flow geometry methods such as Newton's method can be applied.
114 Rheology and laminar flow Flow in circular tube 115

Table 5.4 Bingham plastic, power law and Newtonian formulas for tube flow The special cases of interest are summarized in Table 5.4. Several equations shown
Binghamplastic Powerlaw Newtonian
in the last columns are often called the Hagen-Poiseuille law.
Table 5.5 shows appropriate equations when the flow rate (or average velocity) is
, = KY" '=IJ.Y given. Note the explicit solution available for a Bingham plastic fluid.
Umax =
D(1-rpf,,,
Urn""
Dn
= 2(1 + n)
('w)l/"
K
Dr",
Umax = -
4/L
:rrrYn (-r,,) 11" Example 5.2 Dissipation Rate
q = 8(1 + 3n) K
The power required to pump an incompressible fluid in a straight horizontal conduit

U,vg
Dn
= 2 + (1 + 3n)
,,)1/"
K uavg =
o-;
s; is given by q x ts p, The power per unit volume that goes into viscous losses is the
dissipation rate (Denn, [5]):
u""'" = __ 6__ Um"" 1 + 3n q x 6.p
uavg 3 + 2cp + (P uavg == l+n Dv = :rrD2L .
35(3 + 2cp + qi)3 (1 + 2n)(3 + 5n) 4
CtKE = 54(35 + 5&/1+ 47r/!2) CiKE = 3(1 + 3n)2 Derive an expression for the dissipation rate of a power law fluid flowing in a pipe.

q= 8(1+3n)
1rD3n (D 6.p)
4KL
lin
How does the dissipation rate vary with the diameter if n 0.57 =

NRe = Duavgp Solution


IJ.
We can eliminate either the pressure gradient, the flow rate, or both. From Tables 5.4
16
f=-
N~,

Table 5.5 Wall shear rate, wall stress and pressure drop for different fluids
Binghamplastic Powerlaw Newtonian
o, = [2 3n+4)f-"-1 c:
and 5.5 the following expressions are derived:

3n ) " KqJ.;n] D-(3n+3) = [(1: 3n) (4%) (Hn)/.] D(l+)/'

r = KY" r -!LY

For the specified flow behavior index, the dissipation rate, D., decreases with the 4.5th
power of the diameter, if the flow rate, q is fixed (see first row). Similarly, if the pressure
gradient is kept constant, D. increases with the cubed diameter (see the second row).

6.p
L
4ry
Dr/!
6.: = 23n+2)f-nK ( 1: 3n

xD-(3n+l)qn
r Finally, from the last row the interesting observation is made that D; is constant, if the
nominal Newtonian shear rate is fixed. 0

$._ _8(;_1-=+=-b.:.:_o)
_ bl 5.3.2 Flow Curve
$.
4> = --!...._.._~2
:...:....:...-~ The nominal Newtonian shear rate can be defined for any non-Newtonian fluid as
bl = 2(b;-I/3 + b~/3) (8uavg/ D), but in general it differs from the true wall shear rate. The wall shear rate
bz=l+2bo+b~ for a yield power law fluid depends on both nand <J>
J-
+ 4b-o-+-6'-,b5,....+-4-b~03
-+-b-ri
. 6n3 + lln2 + 6n + 1 (8Uavg)
where y - --- (5.51)
bo = 24IJ.pq
w - 8n3 + 12n2 + 4n - (4n2 + 4n)<J> - 8n2<J>2 - 8n3<J>3 D
1rIJ3r:)' and, thus, the wall stress
tJ.p
L and r/! as above, where t - T +K
6n3 + lln2 + 6n + 1 ) (8u--
It avg)"
w - Y ( 8n3 + 12n2 + 4n - (4n2 + 4n)cp - 8n2<J>2 - 8n3cp3 D
bo = 6IJ.puavg (5.52)
Dry
---- - ._--- -_. - -- -----------_._. ---- ----

116 Rheology and laminar flow Flow in circular tube 117

is not a simple function of the nominal Newtonian shear rate. The apparent viscosity Equation 5.55 tells us that the slope of the log-log straight line is the flow behavior
at the wall for a yield power law fluid can be given as index, n, the same as in the case of the true rheological curve. (No base is given for
the logarithm since both the ten-based or natural logarithm can be used.) The intercept
r
J1-w = y (5.53) at unit nominal Newtonian shear rate is (1 + 3n /4n)n K, and it is often denoted by
( 6n3 + lln2 + 6n + 1 ) (SUDavg) K p- As seen from Eq. 5.55, the true consistency index, K, can be obtained from the
Sn3 + 12n2 + 4n - (4n2 + 4n) - 8n22 _ 8n33 intercept by applying the relation

The advantage of the nominal Newtonian shear rate is that it can be observed.
The plot of the wall stress as a function of the nominal Newtonian shear rate is the K =
4n
( 1+3n
)n Kp. (5.56)
flow curve. The difference between the true rheological curve and the flow curve is
shown in Figure 5.12.
Unfortunately for other fluids, the flow and rheological curves differ more signif-
The flow curve coincides with the true rheological curve only if the fluid is icantly. Formally, it is possible to write the wall stress vs nominal Newtonian shear
Newtonian. The flow curve and the true rheological curve have still similar shapes rate relationship for the general fluid in the same form as for the power law fluid:
in the case of a power law fluid. Indeed, on log-log paper both curves will be straight
lines. From Table 5.5 we have
'w -
_ K' (8Uavg)n'
PD' (5.57)
log(,w) = log(K) + n [log ( ~1+3n) + log (8U)]
;Vg . (5.54)
In this case, however, n' is defined as the log-log slope of the flow curve
or, in a simpler form

(5.5S)
(5.55)

10'
and K'P is defined to give the measured wall stress if substituted into Eq. 5.57.
Naturally, both parameters are varying along the flow curve except for a power
law fluid.
From Eq. 5.51 we see that the wall shear stress is a unique function of the nominal
Newtonian shear rate for a rather complex fluid such as the yield power law fluid.
One of the main results of rheology is that a similar correspondence exists for any
'Cij'
!e:_ constitutive equation. Knowing the flow curve obtained in a tube of any diameter,
p the information is sufficient to determine laminar pressure drops in (other) pipes
.s provided we do not leave the range of nominal Newtonian shear rates corresponding
<I)
10'
~
u; to the experiments. Moreover, knowing the flow curve (or a part of it) we can
~
(I)
construct the true rheological curve (or a part of it). This exercise is based on the
.c:
en Rabinowitsch-Mooney equation [6,7]:

. 1 [
Yw = rrD3 24q +8 L (~P) dq j
d ( ~:)
(5.59)

10"
10-' 10" 10" 1~ 1~ 103
which states that the actual wall shear rate can be obtained from measurable quan-
Shear rate, Y (1/s) tities: the flow rate and its semi-logarithmic derivative with respect to the pressure
Figure 5.12 Rheological curve and tube flow curve gradient.
-----.- _-------------

118 Rheology and laminar flow Flow in circular tube 119

It is tempting to use Eq. 5.57 to characterize a general fluid. The well known form is applied to "adjust" the pipe value. Note that in this book, subscript p is used if the
of Eq. 5.59, coefficient is defined in terms of nominal Newtonian shear rate for pipe flow. The
. =
Yw
(1 +4n'3n') 8uD'
avg (5.60)
meaning of the prime is somewhat different. It indicates an experimentally derived
value valid only for a portion of the flow curve. Other conventions are also used
in the literature. Some authors do not use the subscript p, and other authors might
may suggest the misleading impression that every fluid behaves like a power law prefer the superscript prime for the true power law parameters. The reader should
fluid. This is not true. If we define N as the similar derivative of the true rheological be cautious if data denoted by K' and n' are given without explanation.
curve,
d(ln r)
N=-- (5.61) 5.3.3 Equivalent Newtonian Viscosity for Tube Flow
d(ln y)'
By virtue of the Hagen-Poiseuille law for tube flow, the equivalent Newtonian
and compare the two derivatives at the same stress value, we obtain the relationship
viscosity for a non-Newtonian fluid is defined as
(Babok and Navratil [8])

n' D2 D.p
(5.62) Jote= ---. (5.65)
N = 1 dn' 32uavg L
1-----
3n' + 1 d(ln <",)
In other words f1.e = <wlY",N. The name apparent viscosity is also common for the
In general the inequality same quantity but we prefer the adjective "equivalent" to avoid confusion with the
_J.
y", -r-
(1 + 3N) 8
4N
Uavg
D (5.63)
geometry independent apparent viscosity defined by Eq. 5.6. The apparent viscosity
calculated at the actual wall shear rate is called wall viscosity, Jot", = <",IY",. Table 5.6
shows the different viscosities and their relation to the generalized Reynolds number,
holds, since for a general fluid the two slopes are different. If, however, there is N~ for tube flow. In fact all these equations are mere restatements of the wall shear
an interval of nominal Newtonian shear rates, in which n' is constant, then the true stress vs. average velocity relationship. The apparent viscosity curve on Figure 5.13
rheological curve has a part in which N is also constant and agrees with n', is a property of the fluid while the equivalent Newtonian viscosity curve is valid
In principle, Eq. 5.59 can be used to convert a flow curve into a true rheo- only for the given geometry. The definition of the equivalent Newtonian viscosity
logical curve. The graphical or approximate numerical differentiation to obtain n' may also be applied to non-laminar flow conditions. This practice, however, is not
should be carried out at every point. Then the true rheological curve (wall stress recommended.
vs. true wall shear rate) can be constructed from point to point. Because of the
uncertainty in graphical and numerical derivative calculation, this procedure is not Table 5.6 Viscosities and Reynolds number for non-Newtonian Fluids (lube
recommended. flow)
The application of computers necessitates models with numerical parameters rather Binghamplastic Powerlaw
than the traditional, graphically presented, curves. This leads to the problem of pattern
r = Kv"
recognition and model identification. If the fluid is Newtonian, the flow curve is a

( ~1 + 3n ) "-1( 8uD )"-1


straight line drawn through the origin. If the fluid obeys the power law, we obtain a avg

straight line when plotting on log-log coordinates. For the Bingham plastic and yield JLw=K
power law fluid, neither type of plot can give a well recognizable pattern. Thus,
the selection of a constitutive equation is not trivial. Once an equation is selected, JLt = 2"-3
(
1+
-- n
3n)" KD1-nun-1avg
a nonlinear parameter estimation procedure can be applied to fit the solution to a
given constitutive equation (e.g. Eq. 5.48) to the measurements. N' _ UavgpD(l - t,b)(3 + 2 + 2) N' _ 4nuavgpD
A further observation on K~ can be made. As defined by Eq. 5.57 it refers to pipe R. - JLw Re - (1+3n)JLw
flow. Often the transformation ' _ uavgpD
N R ---
, J.1.e
, (4n'-- )n' K,
K= (5.64)
1 + 3n' P
120 Rheology and laminar flow Flow in circular tube 121

Example 5.3 A Limiting Property of Power Law Fluids


Show that if the nominal Newtonian shear rate is l/s, the equivalent Newtonian
viscosity of a power law fluid cannot be greater than 1.15 [sl-"]K.

~ 10' Solution
'"
te
e:.. From the definition of the equivalent Newtonian viscosity and using the expression for
:::t
the wall stress in Table 5.5
~
'"o
0
1+3n)"
s'" u, = ( 4;;- Ky~:;:/ = (1+371)"
~ x K X [51-"].
lao
A plot of the function (1 + 3n/4n)" (see Figure 5.14) reveals that the maximum occurs
at 71 = 0.241, where (1 + 371/471)" =
1.15. Thus f-Le::: 1.15 K[sl-n]. 0

Example 5.4 Viscositiesof Bingham Plastic Fluid


10" lao 10' 102 Calculate the equivalent Newtonian viscosity, /le, and the wall viscosity, iJ..w, of a
Y
Shear rate, (l/s) Bingham plastic fluid (ry =
5.4 Pa, f-Lp = 0.21 Pa-s") flowing in a tubing of inner
=
diameter, D 3 em. The average velocity is uavg = 0.025 m/s.
Figure S.13 Apparent and equivalent Newtonian viscosity

1.16 Solution
The key quantity to determine is the ratio of the yield stress to the wall stress, .p. From
1.14 Table 5.5

= 6iJ..puavg = 6 x 0.21 x 0.025


1.12 bo -c:-:-,----,-- = 0.194,
Dry 0.03 x 5.4

s,
1.1 b2 = 1 + 2bo V
+ b5 + 4bo + 6b6 + 4b6 + b~ = 2.44,
~
::!:- 1.08
b! = 2(b2"IJ3 + b~/3) = 4.18.
<,

'2 8(1 + ho)


C') ----;o=---b1
+
~
1.06
,,;r;; = 0.67.
~
~ 2
1.04
The required viscosities are calculated from the equations in Table 5.6:

1.02 Dry 0.03 x 5.4


f-Le = 8uavg = 8 x 0.025 x 0.67 = 1.21 Pa- s,
and
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
(3 - if} _.p2 - )D . (3 - 0.673 - 0.672 - 0.67) x 0.03
Flow behavior index, n f-Lw= f-Lp + Ty = 0.21 + 5.4..;__--------.;.._--
24uavg 24 x 0.025
Figure 5.14 Plot of the [(1 +3n)/(4n)]" function
=0.64Pa,s.D
-----~-----------~--
.. --~.-
.. -.-~.. ---..- ..--- .. _..-----------

122 Rheology and laminar flow Flow in other cross sections 123

5.4 Flow in Other Cross Sections For Newtonian fluids a closed form solution can be obtained, while for the Bingham
plastic and power law fluids, good approximations are available (Whittaker [3];
In hydraulic fracturing we have to deal with flow through channels of more complex Savins [9]). The approximate formulas in Table 5.7 give less than 1% error for
shapes. The annulus between two concentric cylinders is of some interest because' w > 0.5 and n > 0.2.
fracturing fluid is often pumped down through the annulus between the tube and The product f x NRE varies from 16 (if w -+ 0) to 24 (if co -+ 1).
casing of a hydrocarbon well. For the description of the flow in the fracture itself a
channel of elliptical cross section and infinite aspect ratio can be postulated.
Example 5.5 Annulus Design
5.4.1 Flow in Annulus =
Water (p = 1000 kg/rn", JL 1 m Pa- s) flow is driven by gravity (g 9.81 mls2) in a =
=
vertical annulus of larger diameter, D2 1 em. Calculate the smaller diameter to assure
For the flow geometry shown in Figure 5.15 the characteristic length is
uavg = 1 m/s,

(5.66)
Solution
where D2 is the larger and DJ is the smaller diameter. The velocity profile depends

( 6.p) pg = 9.81 kPa/m.


on the ratio =
(5.67) L friction

From Table 5.7

Uavg=- ~ (6.p)
- ( 1+w 2
+--
1-W2)
32JL L friction In w

i.e.
1 - w2)
1 = 32 001
. 2
x 0.001
x 9.81 x lif
(~
1 + or: + -1--
nw
.

Using trial and error (or some modern computer software) it is readily determined that
= =
the solution is (I) 0.78. Since, co DdO.Ol m, the inner diameter is DJ = 7.8 mm. 0

5.4.2 Flow in Elliptic Cross Section


Figure 5.15 Schematic of flow geometry in annulus
If we denote the smaller diameter by DJ and the larger diameter by D2 (see
Figure 5.16) the cross sectional area is given by
Table 5.7 Annulus flow
Bingham plastic Power law Newtonian
(approximate) (approximate)

48J.LpUavg L 12u.vg [ 4K L ] lIn


D~(l- w2) 6.p Dz(l - w) Dz(! - w) 6.p with perimeter
(5.69)
= 1- 6ry L 3n
Dz(l- w) 6.p = 1+2n
where (m) is the complete elliptic integral of the second kind and m = 1 - DVD~.
+0.5 [ 4ry ~] 3
(The above equation gives the perimeter ofa circle for m = 0 since (0) = rr/2.)
D2(1- w) /:,p
Applying the definition, the characteristic length is
124 Rheology and laminar flow Flow in other cross sections 125

Figure 5.16 Schematic of elliptic cross section flow geometry Figure 5.17 Schematic of limiting ellipsoid flow geometry

1= TrDl . where
(5.70) rrwQuavgP
2E(m) NRe = -----"'- (5.76)
211-
For Newtonian fluids the average velocity is a simple expression of the pressure
The wall stress is not constant along the perimeter and the definition of the friction
gradient (Happel and Brenner [10]):
factor refers to the average wall stress, denoted by rw' Thus,

(5.71) (5.77)

and the same relationship rewritten in friction factor vs Reynolds number form is An important rearrangement of Eqs. 5.73 and 5.76 is

x N = 2rr2(2 - m)
(5.72)
_Tw=J.L (2TrU
--- avg) , (5.78)
f Re [E(mW' Wo

showing that the average wall shear stress is


5.4.3 Limiting Ellipsoid Cross Section

For hydraulic fracturing applications, the case with infinite aspect ratio, i.e. m = 1, Yw= (2TrU
-r-
--- avg) . (5.79)
Wo
is of great importance. To see the analogy with slot flow, we use the notation Wo for
the smaller diameter (i.e. the maximum fracture width) and h for the larger diameter Solutions for non-Newtonian fluids are not available for the limiting ellipsoid cross
(fracture height) as indicated in Figure 5.17. Substituting E(l) = 1 into the above section. The importance of the power law equation for fracturing fluids necessitates
equations we obtain the expressions for Newtonian flow in limiting ellipsoid cross a formula similar to the one available for slot flow. Attempts have been made to
section suggest a reasonable solution to this problem. In the petroleum engineering literature
U __ w20 Llp
_ the reasoning of Perkins and Kern [11] is accepted. They compared the pressure
avg - 16J.LL ' (5.73) gradient for the flow in a slot with width wand in a limiting ellipsoid cross section
with maximum width Wo = w, provided that the flow rate of the Newtonian fluid is
i.e. the same. The ellipsoid pressure gradient 16/(31l') times larger than the slot one as
Llp 64J.Lq seen from Eq. 5.74 and from Table 5.2. Perkins and Kern assumed that the same
-=-- (5.74)
L rrhw~' relationship holds for power law fluids and therefore

and
f X NRe = 2rr2, (5.75)
Llp
-
2n+5
= --
(1 + 2n)
---
n
Kq h
n -n
w
-(2n+l)
, (5.80)
L 3rr n
-----------_._" .._--------_---------_. ------_. __ .._._ - ..

126 Rheology and laminar flow Flow in other cross sections 127

i.e. Table 5.9 Limiting ellipsoid flow of power law fluids according
~p = 25-n3-1
L it
n-l (1 + n
2n) k n -(2n+l)
uavgw . (5.81)
to Eq. 5.87

6.P -23+" -1 [l+(rr-l)n]n K n -1"+1)


T - it n UavgWo .
The explicit expression of the (average) wall shear rate can be readily recon-
structed from Eq. 5.81:
T
/).P
=
(8)"+1 [l+(rr-l)n]n
-; n
K
q
"h-" -(z,,+I)
Wo

(5.82)

In-I( )"-1
Such a wall shear stress dependence on the flow behavior index seems to be unusual
because of the exponent lin. (At small n the wall shear rate tends to infinity much
faster than for a slot flow.)
f.l.w =K
[ 1 + (1T -
itn.
l)n 2Jtuavg
Wo

Kozicki and Tin [12J suggest another formula for limiting ellipsoid flow

/::;.p = 8K (0.3048 + o.9253n)n (16Uavg )n (5.83)


L l!WO n lrWo

from which one can reveal its origin:

~p = 8K (1 + 3n ) (2rruavg)
n n ,
(5.84)
from which
~p = ~ [1+ (rr - l)n]n (2lTUavg)n
L lrWo 4n lTWO (5.87)
L lrWo rrn Wo
i.e. the wall stress is assumed to be
can be readily obtained. While for Newtonian fluids (n = 1) all the listed expressions
(Eqs. 5.80, 5.83, 5.87) give the known pressure gradient, the limiting behavior at low
(5_85) flow behavior indexes is different as shown in Table 5.8.
From Table 5.8 it is obvious that only Eq. 5.87 results in the desired expression for
In other words, the same dependence of the wall shear rate on the flow behavior near zero flow behavior index. Some other consequences of the suggested equation
index is assumed for limiting ellipsoid flow that has been found for tube flow. The are given in Table 5.9.
analogy is, however, not satisfactory if n is nearly zero as seen from Table 5.8.
A closer look at the "small n behavior" of the wall shear rate both for pipe and
Example 5.6 Flow in Fracture Assuming Limiting Ellipsoid and Slot
slot flow (see the first two columns of Table 5.8) reveals that (if the problem has a
solution at all) the limiting ellipsoid wall shearshould behave as avglwon. Based 2u Cross Section
on this analogy in this work we suggest the following wall shear rate expression A fracturing fluid is reported to have properties K' =
0.004 lbf-s" /ft2, n' = 0.5 and
p = 65.6 lbm/fr'. For a limiting ellipsoid cross section of maximum width, Wo 0.3 =
in, calculate the Reynolds number, the friction factor and the pressure gradient. The
(5.86)
=
average linear velocity is uavg 2 ftJmin. Compare the pressure gradient with flow in
a slot of width w = 0.3 in.
Table 5.8 Wall shear rates at near zero flow behavior index for
power law fluid Solution
Tube Slol Limiting Limiting Limiting We have to decide how to interpret the given rheological parameters. In the absence of
ellipsoid ellipsoid ellipsoid additional information the best we can do is to assume that K' is an adjusted (geom-
(Eq. 5.80) (Eq. 5.83) (Eq. 5.87)
etry independent) power law consistency index. Therefore, we use power law with
2uavg 2u,vg ( ~) lIn JrUavg 1tUavg 2uavg parameters K' and n = n', Using SI units the data are
Dn wn 3 2wOn 2won wOn
K = 0.004 Ibf-s" /ft2 = 0.192 Pa s",
......... _-_.- .._----_._._--------------------------:::------------

128 Rheology and laminar flow References 129

n =0.5, 10. Happel, J., Brenner H.: Low Reynolds Number Hydrodynamics, Prentice Hall, Engle-
wood Cliffs, N.J., 1965.
Wo = 0.00762 m, 11. Perkins T.K. Jr. and Kern, L. R.: Width of Hydraulic Fracture, JPT, (Sept.), 935-49,
uavg = 0.102 m/s, 1961; Trans. AIME, 222, 1961.
12. Kozicki, W. and Tin, c.: Parametric modeling of Flow Geometries in Non-Newtonian
P ;: 1050 kg/nr'. Flows, in Encyclopedia of Fluid Mechanics, (Cheremisinoff, units ed.), Gulf, Houston,
Vol 7, pp. 199-252, 1986.
From Table 5.9

and

The pressure gradient can be determined either directly from Eq. 5.86 or from

t:>.P _ 8 I 2 8 1 2
-L - -f
1lWo
'j_UavgP = :rr x 0.00762 X 0.371 x 2" x 0.102 x 1050

= 673 Palm (0.0297 psi/ft).

The pressure gradient assuming slot flow is obtained from the expression contained in
Table 5.2:

sr = 2"+1 Kw-(n+l)
L (1 +2n)"
-n- U~vg = 521 Pa/m (0.0231 psi/ft).

Thus, slot flow at the same average velocity results in less pressure drop than limiting
ellipsoid flow if the width of the slot is equal to the maximum width of the ellipsoid. 0

References

1. Metzner, A.B., Flow of Non-Newtonian Fluids, in Handbook of Fluid Mechanics,


V.L. Streeter (ed.), McGraw-Hili, New York, 1961.
2. Ullmann's Encyclopedia of Industrial Chemistry (ed. Barbara Elvers) Vol. B,1. Funda-
mentals of Chemical Engineering, Chs 4 and 5, WCH Weinheim, FRG, 1990.
3. Whittaker, A. (ed.): Theory and Application of Drilling Fluid Hydraulics, IHRDC,
Boston, MA, 1985.
4. Chakrabarti, S., Seidl, B., Vorwerk, J. and Brunn, P.O.: The Rheology of Hydroxypropy-
Iguar (HPG) Solutions and its Influence on the Flow Through a Porous Medium and
Turbulent Tube Flow, Respectively (Part 1), RheologicaActa, 30, 114-123, 1991.
5. Denn M. M.: Process Fluid Mechanics, Prentice Hall, Englewood Cliffs, N.J., 1980.
6. Rabinowitsch, B. Z. Phys. Chem., 1, 145A, 1929.
7. Mooney, M., 1. Rheo!., 2,210, 1930.
8. Bobok, E. and Navratil, L.: Folyasi gorbek turbulens tartomanyanak meghatarozasa cso-
viszkozirneterrel, (in Hungarian) Koolaj esFoldgtiz, 24, 135-144, 1991.
9. Savins, J.G.: Generalized Newtonian (Pseudoplastic Flow in Stationary Pipes and
Annuli, Pet. Trans. AIME, 213, 1958.
6
NON-LAMINAR FLOW
AND SOLIDS TRANSPORT

The transitional and turbulent behavior of fracturing fluids has not been studied
adequately and a similar statement is even more true for the transport of solids by
these fluids. The general approach is to find the essential variables and/or their combi-
nations and to present the relationship between them in graphical and/or algebraic
form. This chapter is devoted to relations relevant to the flow and solids transport in
the fracture and in the well tubulars.

6.1 Non-laminar Flow


6.1.1 Newtonian Fluid
Transitional and turbulent flow has been studied extensively for simple geometries
and fluids. For a wide class of Newtonian fluids the friction factor, f, in a tube
is determined by the Reynolds number, N Re, and the relative roughness of the
wall, e. The log-log plot of f versus NRe, shown in Figure 6.1, is the well known
Moody diagram [1] generated with an explicit relationship presented below. The
Colebrook- White [2] equation

1 ( e 1.255) (6.1)
.Jl = -410g10 3.7 + NRe.Jl
provides a basis to calculate the friction factor numerically. The inherently iterative
method can be substituted by an explicit approximation offered by Chen [3]

_1__
.Jl-
-410
glO
{..!._
3.7
_ 5.0452
NRe
10
glO
[81.1098 (7.149)0.8981]}
2.8257 + NRe . (6.2)

The relative roughness can be considered as a value on an empirical scale, origi-


nally defined by Nikuradze [4] who created artificial roughness by gluing sand to the
inner wall of smooth pipes. In practice, the relative roughness is either assumed or
Non-laminar flow 133
132 Non-laminar flow and solids transport

where
10"'
, = K'
f.l-e P
(8 )
uavg
D
n'-l
(6.5)

Both f.l-~ and NRe reduce to the corresponding power law values (see Tables 5.4
and 5.6) if the fluid obeys the power law. The basic assumption is that from the point
.....
of view of the turbulent behavior any fluid can be replaced hypothetically by a power
..: ~
law fluid. The replacing power law fluid is selected to give locally similar pipe flow
~ O.OS
S 0.04
0.03
behavior in the corresponding shear rate range but under laminar conditions. In this
;;0
c::; 0.02 !!!. formalism the use of the symbol K~ underlines the practical character of the approach.
1~ ~ 0.015
0.Q1
~ Indeed, only observable variables are involved in the transfer of information from
0.006
~ 0.004
(1)
the laminar to turbulent regime. The specific use of the power law is not an absolute
CD ~ I- 0.002
;;0
0
.s
c::;
~ 0.001
c::;
<C necessity and other authors prefer to use a generalized Reynolds number derived
c::;
as
LL
IIiIII 0.0006
0.0002
:T
::J
(I)
from other models (but still obeying fNRe =
16, if the flow is laminar.) Such a
I'IIItiiII[::: <J)
construction is given in Table 5.6 for the Bingham plastic fluid.
0.0001 !"
I I' 0.00005 M
For a Bingham plastic fluid the above "generalized power law" treatment can be
0.00001
compared numerically to more accurate turbulent flow description. Hanks [6] shows
0.000001
that the deviation could be significant and, therefore, the generalized power law
1(t3 concept should be used with caution.
1()2 10" 10~ 106 The Dodge-Metzner approach has been extended for rough pipes by Szilas
et al. [7]. Their BNS equation is rewritten here in terms of the Fanning friction
Reynolds nurnosr, NRe
factor:
Figure 6.1 Frictionfactor vs. Reynoldsnumber [lOY
- 1 --410 g10 (Re,)1/n' f(2-n')/(2n') e ] (6.6)
J7 - + --
3.715 '
determined experimentally from the Colebrook-White equation using a fluid (e.g.,
water) known to obey it. where
y = 0.8295 + 1.405 _ 1.5111/n' (0.3535 + 1.061) . (6.7)
n' n'
6.1.2 GeneralFluid Naturally, the BNS equation reduces to the Colebrook-White equation for n' = 1.
For a wide range of non-Newtonian fluids flowing in a smooth pipe the approach of
Dodge and Metzner [5] can be applied: Example 6.1 Turbulent Flow of a "Generalized" Fluid

_1 4_ ' (2-n')/2 _ ~ (6.3) A fracturing fluid (K~ = 0.012 lbf-s" /ft2, n' = 004, p = 65.5 Ibm/fr') is pumped
VI - (n,)O.75 loglO(NRef ) (n')1.2 . through D = 2.259 in. tubing. The flow rate is q = 20 bpm. Neglecting possible drag
reduction and assuming a relative roughness of 0.001, calculate the frictional pressure
Here the meanings of NRe and n' require more clarification. First, the nominal Newto- gradient.
nian shear rate is calculated. Note that it is nominal not only because the fluid is
not Newtonian but also because, in spite of the different velocity profile due to
Solution
turbulence, the calculation involves a hypothetical laminar velocity profile. Second,
it is also assumed that the laminar flow curve in the vicinity of the nominal Newto- Before the actual calculations we have to interpret the given data carefully. Since the
nian shear rate can be described by Eq. 5.57 introduced in Chapter 5. The Reynolds subscript p is given explicitly, we use this value as a geometry dependent(pipe) consis-
number is calculated from tency index: K~ = 0.012 Ibf-s"'/ft2 = 0.575 Pa s"'. Note that the adjusted (geometry
independent)value would be K' = 0.216 Pa- sn', but we do not need it, because Eq. 6.4
N
, uavgpD
- --
gl-n'
- -____;~--
pDn' u~;t (6.4)
can be used directly.The Reynolds numberis obtainedfrom the diameter,density,linear
Re - /I'r: - K'p velocity and the pipe consistency index:
----._,,_ .. __ ..

134 Non-teminsr flow and solids transport Non-laminar flow 135

D = 0.0574 m,

p = 1050 kg/m",
q
II = --W = 20.5 m/s,
d2_
4
81-nJ 2-n' tr'
N'
R. -
- u
K'
p = 2.54 X 105.
p

The solution of the BNS equation (Eqs, 6.6) for n =


0.4, e = 0.001, N n. 2.54 x 105 =
is f = 0.00490.
If we neglect that the Reynolds number has been obtained from a power law, we
could use the Colebrook-White equation (Eq, 6.1). Its solution for e == 0.001, Nne =
2.54 x 105, would yield f =
0.00519. The same value could be obtained directly from
the Chen approximation (Eq. 6.2).
The pressure drop is calculated from the definition of the friction factor:
Do 4.
: = D [0.00490 X Opu2) J = 7.53 x lit Pa/m (3.33 psi/ft), MDRA
11111111
which is 7.6 times larger than the pressure gradient due to gravity. 0
v II IIII
101 10 105 107
6.1.3 Drag Reduction Wall Reynolds.number, NRe,w

Most fracturing fluids which are polymer solutions exhibit drag reduction in turbulent Figure 6.2 MaximumDrag Reduction Asymptote
flow. The term drag reduction can be applied if a certain fluid shows Newtonian
behavior in laminar flow, but for higher Reynolds numbers the friction factor is below to Eq. 6.8 is available (Denn [9]):
the value calculated from the Colebrook-White equation. As a first approximation, In(!) = 28.135 + (-29.379 + (8.2405 - 0.86227x)x)x,
a similar definition might be applied to a general fluid: drag reduction means that
where x = In[ln(NRe)]. (6.9)
the turbulent friction factor is significantly less than the one calculated from the
Dodge-Metzner relation. A somewhat simplified view of drag reduction envisions the Most fracturing fluids have one or two orders of magnitude higher polymer
long polymer molecules dampening the instabilities and hence reducing the turbulent concentrations than the solutions studied by early investigators of the drag reduction
energy dissipation. As a consequence, the transition from laminar to turbulent regime phenomenon. Therefore, it is natural .ro assume that maximum drag reduction is
is gradual. in effect for these shear thinning fluids. Nevertheless, the question is still open
The concept of drag reduction was originally used to indicate the reduction of whether the MDRA can be applied numerically to shear thinning fluids. The principal
friction pressure when small quantities of high molecular weight polymers were reason of the uncertainty is that one has to decide which Reynolds number to use
added to turbulent water flow (Virk, [8]). When the addition of the polymer is so in conjunction with the above equation. Virk [8] criticized the early practice to use
limited that the water viscosity remains constant, the comparison can be made at the solvent Reynolds number (i.e., water Reynolds number for water based polymer
the same flow conditions characterized by the solvent Reynolds number. It has been solutions) and stressed the importance to correct for the shear thinning. Moreover,
observed that the addition of polymer decreases the friction factor only to a limiting he stated explicitly that "a Reynolds number formed with the apparent wall viscosity
value. For dilute polymer solutions Virk [8] established the Maximum drag reduction seems preferable to the generalized power-law Reynolds number". The wall Reynolds
asymptote (MDRA) number is calculated with the wall viscosity
Duavgp
(6.8) NRew= ---. (6.10)
, tJ-w
providing the lowest possible friction factor at a given Reynolds number. The plot Explicit expressions for the wall viscosity of the Bingham plastic and power law
of the MDRA is shown in Figure 6.2. A sufficiently accurate explicit approximation fluids were given in Table 5.6. Note that not even for a power law fluid the wall
136 Nonlaminar flow and solids transport Nonlaminar flow 137

Reynolds number and the generalized Reynolds have the same values: The solution of Eq. 6.8 is f =
0.565 x IfY. (The explicit Eq, 6.9 yields virtually the
same value: f = 0.563 x 10-3.) The corresponding pressure gradient is

NRe.w == ( 4;-
1+ 3n) ,NRe (6.11) ~p 4
L = '0[0.00565 I 0
x (zpu-)] = 0.868 4
x 10 Palm (0.384 psi/ft).

The definition of the wall Reynolds number can be applied to any other constitutive
The calculated frictional pressure drop is less than the pressure increase due to
equation, as well. For example, for the yield power law first a fictitious laminar wall gravity and much smaller than the one calculated in Example 6.1. The above result
stress (or pressure drop) has to be determined from Eq.5.48 and then the wall should be treated with some caution because of the still open questions whether the
viscosity can be calculated dividing the wall stress by the wall shear rate (Eq. 5.51). MDRA can be applied and whether the wall Reynolds number should be preferred.
Knowing the wall viscosity we can proceed to obtain the wall Reynolds number. Nevertheless, from the results of Examples 6.1 and 6.2 it should be clear that the drag
From the wall Reynolds number the friction factor and the turbulent pressure gradient reduction phenomenon cannot be neglected. 0
can be calculated according to Eqs. 6.10 and 6.9, respectively.
There have been several interesting attempts to characterize the turbulent behavior
of viscoelastic fluids involving another dimensionless group, the Deborah number 6.1.4 Turbulent Flow in Other Geometries
(Dem [9J). The somewhat general definition of the Deborah number is the ratio
Turbulent flow in other geometries such as a flow channel with an elliptic cross
of the characteristic relaxation time of the fluid to the characteristic process time.
section has not been investigated extensively. A general rule of thumb is suggested
When the Deborah number is small, the fluid behaves like a purely viscous liquid.
as follows. The relevant Reynolds number for the given geometry is determined and
When the Deborah number is large, the elastic property of the fluid dominates
then a friction factor for tube is obtained according to one of the above methods.
its behavior.
The friction factor is modified according to the ratio (f x N Re) /16 valid for the
It is not straightforward to translate the concept into a readily observable quantity
given geometry in the case of Newtonian flow. For slot geometry the ratio is ~
and therefore the Deborah number is calculated extremely differently by different
authors. Studying the flow behavior of a fracturing fluid (HPG) Chakrabarti et al, [lOJ and for limiting ellipsoid cross section it is 2rr2/16. Finally, the pressure gradient is
suggested that the MDRA might be the N De -+ 00 limit of an f vs (N Re, N De) calculated from the friction factor.
correlation. Analyzing the friction factors for the same fluid Keck [11] arrived at It is usually accepted that the kinetic energy correction factor, aKE, can be taken
as unity for turbulent flow in any geometry since the velocity profile is very fiat in
a conclusion that the Deborah number should be in the range 22 to 30 for a
0.48% HPG solution and in this range the friction factors lie considerably below comparison to the laminar profile.
the MDRA.
Example 6.3 Turbulent Flow in Ellipsoid Cross Section
Example 6.2 Pressure Drop from Drag-reduction Correlation
Water (IL = 1 cp, p = 62.4 lbm/fr") is pumped into a two wing fracture as "prepad".
Consider the same conditions as in Example 6.1. The viscoelastic fracturing fluid (K' = Assume a limiting elliptic cross section with height 60 ft and maximum width Wo =
0.4 in. Calculate the pressure gradient if the total flow rate is 20 bpm. Use reasonable
0.012 lbf-s" /ft2, n' = 0.4, p = 65.5 Jbm/fr') is pumped through D = 2.259 in. tubing.
assumptions concerning the roughness of the fracture surface.
The flow rate is q = 20 bpm. Neglecting pipe roughness but taking into account the
drag reduction phenomenon calculate the frictional pressure gradient.
Solution
Solution The data in SI units are:
In Example 6.1 we found that the generalized Reynolds number is q = 20 bpm = 0.0534 m3/s,

N' -
81-n' 2-n'
u
TVI'
p,-, = 2.54 x lOs
h = 60 ft = 18.3 m,
Re - K' Wo = 0.4 in = 0.0102 m,
p

From Eq. 6.10 the wall Reynolds number is IL = 1 cP = 0.001 Pa s,


p = 62.4 lbm/tt? = 1000 kg/m".
N . = (1 +4n'3n')
Re.w
r
N R. _
-
~.5
3.50 x liT.
.----~-------.--.- ...
----------=

138 Non-laminar flow and solids transport Solids transport 139

Table 6.1

e f i [Palm) !J.p [psilft]


L
0.1 0.0329 142 0.00630
0.01 0.0159 691 0.00306
0.001 0.0136 59.0 0.00261

Since the flow is equally divided between the two fracture wings, the linear velocity is
q
U = -h 2 = 0.182 m/s,
won
4
Proppant bed
The resulting Reynolds number is (see Eq. 5.76)

N = n x 0.0102 x 0.182 x 1000 _


Re 2 x 0.001 - 3040.
Figure 6.3 Components of proppant velocity
Thus, turbulent flow regime can be assumed. The friction factor is determined (assuming
an appropriate relative roughness) from Eq. 6.2 taking into account the factor 2n2/16: a part of the proppant bed. The equilibrium height concept is a suitable means to

f _2n
- -2
16
[
-41oglQ
{e--
3.706
5.0452
- --loa
NRe
l.1098
--
010 2.8257
[e
+ (7.149)O.S981]}]-2
--
3040
describe proppant transport under such conditions (Babcock et al. [15]).

6.2.1 Settling of an Individual Sphere


The corresponding pressure gradient is obtained by applying the definition of the friction
factor (Eq. 5.1l) and the relation between pressure drop and wall stress (Eq. 5.77): According to Stokes' law the drag on a sphere of diameter d p moving steadily with
velocity u through an unbounded fluid is given by
tJ.p _ ( 8 ) 1 2
L - tt x 0.0102 (f)(2:0.182 x 1000).
(6.12)
For an order of magnitude analysis it is reasonable to assume that the relative roughness
is minimum 0.001 and maximum 0.1. The results are given in Table 6.1. At particle Reynolds number
The interested reader may compare these results with the measurements of
Warpinski [12], who found that observed in situ pressure losses in a fracture were
from 1.4 to 3.1 times higher than those calculated neglecting the roughness of the
(6.13)
walls. 0
less than 0.01 the drag specified by Stokes' law is only 2 percent less than a value
based on rigorous small Reynolds number expansion of the Navier -Stokes equations
6.2 Solids Transport [9]. In engineering practice Stokes' law is considered valid if N Re.p < 1. The drag
on a sphere can be characterized through the drag coefficient, CD, defined to relate
One of the main functions of the fracturing fluid is to transport proppant. The
proppant carrying capacity of the fluid is limited by gravitational settling. Figure 6.3 the induced stress (force divided by the projected area of the sphere) with kinetic
shows a possible trajectory of the proppant particle in the fracture. A simple vectorial energy per unit volume
view of the horizontal and vertical velocity component suggests the requirement that 4F ) 1 2
the settling velocity should be less than the horizontal transport velocity multiplied ( d~Jr = CD X CzPfU ). (6.14)

by the ratio of height to half-length. This requirement has led to the application of
high-viscosity fracturing fluids (Daneshy [13]; Novotny [14]). Combining the above three equations we arrive at another form of Stokes' law:
If proppant settling is significant a bank of solids is formed leaving less free space
for flow of the slurry. The resulting larger horizontal velocities lead to remobilizing C DN Re.p = 24. (6.15)
==== ===~================ ..__--=======-------
.. - - - - ------

140 Non-laminar flow and solids transport


Solids transport 141

At higher Reynolds numbers the inertial forces become important and the above
product increases. For engineering computations Dellavalle's [16] correlation dnp+1 fl. pg ] lin
u = (6.22)
o [ 18Kf(n)

CD = (0.63+ ~)2
vNRe,p
(6.16) If the power law Reynolds number is defined accordin~ to

N' d"u
_r_P
2-"pf
_
gives sufficient accuracy for a very wide range of Reynolds numbers, 0 < NRe.p < Reip = K (6.23)
2 X 105
then Eq. 6.22 can be rewritten as

Terminal Velocity CDN~e.p = 24f(n). (6.24)

Care should be taken if the resulting terminal velocity corresponds to a shear


A particle whose density is fl.p = p p - pf higher than the corresponding fluid rate that is considerably less than the one at which the power law parameters were
density is accelerated by the gravitational force. With increasing velocity, however, determined. In such case the power law extrapolation overestimates the apparent
the drag force on the sphere acting against the gravitational force is also increased. viscosity. As a result, the predicted terminal velocity might be underestimated as
The falling particle soon reaches an equilibrium when the two forces are balanced: demonstrated by Roodhart [18J.
7l'd~fl.pg
A convenient empirical approach to present the results of settling experiments
37l'fJ,dpuo = 6 ' (6.17) was suggested by Shah [19]. He found that the plot of the dimensionless group

from which the terminal velocity is


V C~-II(NRe.p)2 versus the particle Reynolds number, N~e.p i~ a unique ~urve for a
given flow parameter, n'; and presented the curves both graphically and In the form
d~fl.pg of simple correlating equations.
Uo=--- (6.18)
181
The more general relation equivalent to Eq. 6.18 in the Stokes regime but valid for 6.2.2 Effect of Shear Rate Induced by Flow
any Reynolds number is
For hydraulic fracturing applications we are interested in settling under the additional
(6.19) shear induced by the flow. The flow induced shear results in further changes of
the draa force if the fluid is non-Newtonian. To account for the thinning effect
Novotny [14] suggested the use of Stokes' law with an apparent viscosity calculated
The terminal velocity corresponds to idealized conditions (infinite fluid).
at an "effective shear rate"
Nevertheless, it serves as a suitable reference velocity for describing more realistic
situations.
y = [(;:r + (Ydf/ 2 (6.25)

Non-Newtonian Fluid
where Yt is the shear rate which would be generated by the fluid flow only. Using a
Results of the drag force acting on a moving sphere are available for power law parallel plate approximation the shear stress induced by the fluid flow can. be obtained
fluids. Acharya [17] obtained from the constitutive equation (see Table 5.2). Since Yl depends on the distance from
the wall, the effective shear rate is higher in the vicinity of the wall than in the center.

F = 37l'd~K (:J 11 fen), (6.20)


Assuming power law behavior the equation becomes

where the function f (n) is given by

fen) = 3(3n-3)/2 [33n5 - 63n4 -lln3 + 97n2 + 16n] ;


(6.21)
"0 ~ ~::g{G:)' + [~
21

% Y 1 "[2(l+"~)"'" r } (1-n)/2

(6.26)

4n2(n + l)(n + 2)(2n + 1) We can solve (numerically) the above equation at any y and hence a profil~ of settli~g
velocities can be determined. If only one representative value of the setthng velocity
therefore the power law form of Eq, 6.18 is
is needed, the above equation may be solved at y = w/4.
-----.~--

142 Non-laminar flow and solids transport SOlids transport 143

Dallavalle equation (Eq. 6.16) is still valid in a slurry characterized by the void
fraction of the solids, 4>, if both the Reynolds number and the drag coefficient is
modified according to

N _ (ur/J) ex [ 54>] (dpUr/JPf) (6.27)


Re,p.r/J - Uo P 3(1- 4 JL '
0.00008
~
.s CD,r/J =
UO)2(
( ur/J 1-4> )
1 + 4>1/3
(4F)
d~T(
1 2-1 ,
(2PjUr/J) (6.28)
s 0.00006

where F is still given by Eq. 6.12, understanding that the moving velocity is now ur/).
0.00004 Thus, a terminal velocity calculation should involve Eqs. 6.12, 6.16, 6.27 and 6.28,

0.00002 6.2.4 WallEffects


The terminal velocity decreases (relative to the unbound fluid case) of an individual
0
sphere falling between walls can be described by the wall effect function
0 y w/2
Wall Centerline fw=UOw. (6.29)
UO
Figure 6.4 Settling velocity for Example 6.4
A treatable form of f II' derived theoretically by Faxen (see the detailed treatment
by Happel and Brenner [21]) for a particle located at a distance y = w/4 from the
The velocity field around a falling sphere is disturbed by the presence of walls wall is
and/or other particles, leading to additional effects.
f II' = 1 - 0.652&0 + 0.1475,,/ - 0.131(1)4- 0.0644w5 + ... (6.30)

Example 6.4 Settling VelocityProfile in the Facture where


2dp
Construct a curve representing the terminal velocity of a 20/40 mesh bauxite proppant (J) =-. (6.31)
w
(dp = 6.4 x 10-4 m, Pp = 3700 kg/nr') in a borate crosslinked polymer characterized
by PI = 1010 kg/m", n =
0.68, K = 33.5 Pa- s". Assume fracture width w 0.008 m = Although there is no consensus in the question how to take into account the
and average linear velocity uavg = 33.5 m/s, concentration and wall effects simultaneously, it is reasonable to apply the wall
correction only at the final stage of the calculation in a non-iterative manner.

Solution
Example 6.5 Representative Settling Velocity Involving Slurry
The solution of Eq. 6.26 can be obtained numerically for different distances from the and Wall Effects
wall. The results are plotted in Figure 6.4. The calculated terminal velocity at the
centerline (y = w/2) is only Ui] =
3.4 X 11)-6 m/s, an unrealistically low value. The Repeat the calculations of Example 6.4 for a representative location in the fracture,
terminal velocity at y =
w/4 (uo = 8.6 X 10-5 m/s) can be accepted as a reasonable y = w/4, taking into account slurry and wall effects. Assume that the slurry was
representative value. 0 obtained adding 10 Ibm of proppant per gallon of fluid (10 ppga).

6.2.3 Effect of Slurry Concentration Solution


The presence of other particles causes an additional decrease of the terminal settling Adding 10 Ibm bauxite (pp =
3700 kg/m") to 1 gallon liquid results in fluid volume,
velocity. According to the interesting observation of Bamea and Mizrahi [20] the VI =
1 gallon =
0.00379 m3, and particle volume, V p = 10 Ibm/ Pp = 10 x 0.456/
---------- ...-.-.-----.-----~----
----_--_ ..._-----

144 Non-laminar flow and solids transport References 145

3700 = 0.00123 rrr', Therefore, the void fraction of solids is rp = V p/Wf + V p) = 6.2.5 Agglomeration Effects
0.00123/(0.00123 + 0.00379) 0.245. =
For a terminal velocity, uq" the apparent viscosity is calculated (in accordance with When several settling particles form a cluster, the settling rate might be considerably
Eq. 6.26) from
higher then for individual particles. Since a cluster acts more or less as a large
particle and according to Stokes' law the terminal velocity is proportional to the
square of the particle diameter, the increase of the settling velocity is not surprising.
Clusters may consist of a few particles or more. In hydraulic fracturing the proppant
carrying fluid may be different from the fluid already-present in the fracture and this
33.52 may lead to the limiting case where the proppant and its surrounding fluid behave
(17627 + 2.44 x 106u~)O.16 as a single cluster, moving together in the pool of the other fluid. The phenomenon,
called convection (Cleary and Fonseca [22]), does not lend itself to easy description.
and the Reynolds number is obtained from Eq, 6.27,
Since no qualitative description is yet available, at present the role of the concept is

N Re.p. = G:) exp [- 3(15~ )] (d Pp:: f)


rather to illustrate the "overwhelming complexity of the underlying physics" than to
obtain reliable design guidelines.
= 131.2 x (17627 + 2.44 x 1Q6U!)O.16.

Applying the Dallavalle correlation, Eq. 6.16, the resulting drag coefficient is References

CD.q, = (0.63+ ~)2 Re.p.


1.
2.
Moody, L.W.: Friction Factors for Pipe Flow, Trans. ASME, 66,671, 1944.
Colebrook, C.E.: Turbulent Flow in Pipes with Particular Reference to the Transition
Region between Smooth and Rough Pipe Laws, 1. Inst. Civil Eng., (London), 11, 133,
1939.
-0.63 + 4.8 ) 2
3. Chen, N.H.: An Explicit Equation for Friction Factor in Pipe, Ind. Eng. Chem. Fund.,
(
V131.2x (17627+2.44 x 1Q6U~)016 18, 296-297, 1979.
4. Nikuradze, J.: Stromungsgesetze in rauen Rohren, VDI Forschungsheft, Arb. Ing.-Wes.,
From the definition of the drag coefficient for a particle settling in a slurry Eq. 6.28,
No. 361, 1933.
C .=
D.",
(U)2
Uo
(.2.=..t_)
1 + .p1/3
[4(31T/-(a
d~1T
dpU)] 1 2-1
(7_Pfu)
5. Dodge, D.W. and Metzner, A.B.: Turbulent Flow of non-Newtonian Systems, A.!.Ch.E.
lournal, 5, 189~204, 1959.
6. Hanks, R.W.: Principles of Slurry Pipeline Hydraulics, Encyclopedia of Fluid
we obtain the new approximation of the terminal settling velocity, Mechanics, Cheremisinoff, N. P. (ed.), Vol. 5, pp. 237~240, Gulf, Houston, TX, 1986.

u = (CD.t/>Pfu5) (1+ rpl/3)rp


1-
(_:!_P__)
24fLa
7. Szilas A.P., Bobok, E. Navratil, L.: Determination of Turbulent Pressure Loss of non-
Newtonian oil Flow in Rough Pipes, Rheologica Acta, 20, 486-496, 1981.
8. Virk, P.S.: Drag Reduction Fundamentals, A.l.Ch.E. Journal, 21 (4), 625-656, 1975.
= 1.27 x 10-11 x (17627 -:- 2.44 x 106u;)016 9. Denn M.M., Process Fluid Mechanics, Prentice Hall, Englewood Cliffs, N.J., 1980.
10. Chakrabarti, S., Seidl, B., Vorwerk, J. and Brunn, P.O.: The Rheology of Hydrox-
ypropylguar (HPG) Solutions and its Influence on the Flow through a Porous Medium
x (0.63 + 4.8 ) 2 and Turbulent Tube Flow, Respectively (Part 1) Rheologica Acta, 30, 114-223, 1991.
V 131.2 x (17627 + 2.44 x 106 u~)O 16 11. Keck, R.G.: The Effects of Viscoelasticity on Friction Pressure of Fracturing Fluids,
paper SPE 21860 presented at the Rocky Mountain Regional Meeting and Low-
The solution of the above equation is Uq, = 0.00013 m/s. Permeability Reservoirs Symposium, Denver, Co, Apr. 15-27, 1991.
To include the wall effect, we use Eqs. 6.29 to 6.31 as follows: 12. Warpinski, N.R.: Measurement of Width and Pressure in a Propagating Hydraulic
2dp 0.00064 Fracture, SPEJ, (Feb.), 46-54, 1985.
w= - =2--- =0.16 13. Daneshy, A.A.: Numerical Solution of Sand Transport in Hydraulic Fracturing, 1PT,
w 0.008 '
(Nov.), 135 -240, 1978.
fw =1- 0.6526w + 0.1475w" - 0.131w4 - 0.0644w5 = 0.899. 14. Novotny, E.J.: Proppant Transport, paper SPE 6813, 1977.
u.w = fwut/> = 0.899 x 0.00013 = 0.000117 m/s. 15. Babcock, R.E. Prokop, CL, Kehle, R.O.: Distribution of Propping Agents in Vertical
Fractures, Producers Monthly, (NOV.), 11-18, 1967.
Thus, a reasonable representative value of the terminal settling velocity is 0.12 mm/s, 0 16. Dallavalle, 1.M.: Miocromeritics, 2nd ed., Pitman, London, 1948.
~- . __ .... . --.------.----~------------------------:--.

146 Non-laminarflow and solids transport

17. Acharya, A.: Particle Transport in Viscousand ViscoelasticFracturing Fluids, SPEPE,


(March), 104-110, 1986.
18. Roodhart, L.P.: Proppant Settling in Non-Newtonian Fracturing Fluids, Paper SPE
13905, 1985.

7
19. Shah, S.N: Proppant Settling Correlationsfor non-Newtonian Fluids under Static and
Dynamic Conditions, SPEJ, (April), 164-170, 1982.
20. Barnea, E. and Mizrahi, J.: The Chern.Eng. Journal,S, 111-189, 1973.
21. Happel, J. and Brenner, H.: Low Reynolds Number Hydrodynamics, Prentice-Hall,
Englewood Cliffs, 1965.
22. Geary, M.P. and Fonseca, A., Jr.: ProppantConvectionand Encapsulationin Hydraulic
Fracturing: Practical Implicationsof Computerand Laboratory Simulations,SPE paper
ADVANCED TOPICS OF
24825, 1992.
RHEOLOGY AND
FLUID MECHANICS

In hydraulic fracturing foamed polymer solutions are used widely. They are attractive
propp ant-carrying fluids because of their excellent solids transport carrying property
while they may considerably decrease formation damage due to fracturing, espe-
cially for water sensitive formations. This chapter concentrates on some aspects of
foam flow.

7.1 Foam Rheology


Foams are gas-liquid dispersions with the liquid as the continuous and the gas as
the dispersed phases. The in situ volumetric phase relation is usually characterized
by the quality, I", defined as the ratio of the gas volume to the total foam volume
(Cameron and Prud'homme [1]). Fracturing fluids with qualities lower than 50-55%
have been referred to as energized. Very high-quality foams, above 93-97%, have
the tendency to invert into mist when the liquid becomes the internal and the gas the
external phase.
It is convenient to make an explicit distinction between microflow, where the char-
acteristic size of the space confining the flow is commensurable with the bubble size
(e.g., flow in porous media and in traditional laboratory viscometers) and macroflow,
where the bubble size can be neglected with respect to the characteristic size of the
flow path (e.g., flow in a fracture or well). For the purpose of macroflow description
the foam is described as a homogenous fluid with a "rheology" depending not only
on the shear rate but on additional parameters, first of all on the relative gas content.
(Note that the homogeneous fluid concept presumes a common linear velocity for
both phases at any given location.) It is well established now that the apparent
viscosity increases with increasing foam quality and decreases with increasing shear
rate (Assar and Burley [2]).
"""---.--- ------- ----------

148 Advanced topics of rheology and fluid mechanics Foam rheology 149

7.1.1 Quality Based Correlations Newtonian flow and incompressible non-Newtonian flow along a flow path of
constant cross section obey a certain invariance property: The friction factor (in
A frequent implicit assumption, when using quality-related correlations, is that the other words the Reynolds number) is constant. If we demand the same invariance
pressure (and hence the density of the gas phase, the compressibility of the bubbles) property for the flow of a compressible non-Newtonian fluid, this restricts the form of
has no direct effect on the rheological behavior. the constitutive rheological equation. A constitutive equation, providing the required
Several studies (Blauer et al. [3]; Reidenbach et al. [4]) start from the assump- invariance, is called volume equalized. The general form of shear rate versus shear
tion that a rheological curve, determined at given conditions, undergoes continuous stress functions, satisfying the principle of volume equalization, is
deformation with small perturbations. Once a flow curve is described by a two- or
a three-parameter rheological equation the most influential condition (the quality) is
changed and the transformation of the flow curve is described establishing empirical
~=fVEG) (7.2)

correlations for the parameters of the rheological equation. Thus, Eq. 7.2 states that the volume equalized shear rate is a unique function of the
The above procedure of parametrization can be continued including other factors volume equalized shear stress.
(e.g., by varying the texture via applying different foam generators and/or surfac- In Chapter 5 we saw that a standard flow curve relates wall shear stress and
tant concentrations, fluid and gas composition, system pressure and temperature). nominal Newtonian shear rate. For foams formed from a given gas-liquid pair but
Unfortunately, data reduction with respect to one flow curve, corresponding to a at different qualities and pressures, such a plot looks scattered as seen on Figure 7_1.
prescribed quality, is mathematically ill-conditioned. Several combinations of the By virtue of Eq. 7.2 the wall shear rate versus nominal Newtonian shear rate plot
parameters may reproduce the flow curve within the achievable experimental accu- collapses into one curve if both coordinates are volume equalized. A volume equal-
racy. This is especially true for three-parameter models such as the yield-power ized plot of the data of Figure 7.1 is shown in Figure 7.2. The volume equalized
law. The small variations in the quality and quantity of additives (surfactants, clay flow curve can be described applying any of the known constitutive equations. The
stabilizers and breakers) and in the other conditions (texture and shear history) may parameter estimation procedure is exactly the same as for incompressible fluids,
cause a shift in the flow curve which is then amplified in the numerical values of resulting in the parameters of the volume equalized model.
the parameters. The three-constant constitutive equation
In a detailed investigation of water and HPG-solution foams at elevated pressures
Reidenbach et al. [4] applied a yield-power law formulation. In order to incorporate ;r = !y'vE + K VE (Y)"
-;; (7.3)
the non-Newtonian behavior of the base liquid into the constitutive equation those
authors postulated that the flow behavior index n was identical to that of the base reduces to a yield power law with yield stress t ; = t,.VEC and K = KVEI-11 at any
liquid while the yield stress and the consistency index varied with quality. given specific volume expansion ratio and it is of the form of Eq. 7.2; thus it is
justified to name it volume equalized yield-power law.

7.1.2 Volume Equalized Constitutive Equations


In interpreting large-scale experiments, the rheological behavior of foams of different
qualities could be represented by one curve (Valk6 and Economides [5]) if both the
shear stress and the shear rate were volume equalized.
(ii'
The technique uses the specific volume expansion ratio, e, as the additional param-
eter representing the volumetric relation of the gas and liquid phases. It is defined
~ 10
6
66
""
0

as the ratio of the liquid density (considered to be constant) to the foam density ....!I: 6 r=48 to 50%
(p=517 to 539 kg/m")
(varying along the flow path because of the change in pressure): e r=57106O%
(Jr"421 10447 kg/m")
PI .. r=6710 7oo",
= ~. (7.1) (p=324 to 355 kg/m")
P
The specific volume expansion ratio may vary between unity and a maximum 10 100 1000

determined by the ratio of gas density to the liquid density. It is convenient to refer
to a quantity divided by e as volume equalized.
The principle of volume equalization [5] is based on an observation concerning Figure 7.1 Shear stress vs. shear rate data for foam flow of different qualities, pressures in pipes
non-Newtonian compressible flow. Incompressible Newtonian flow, compressible of different diameters (after f7])
150 Advanced topics of rheology and fluid mechanics Foam rheology 151

variables are determined by numerically solving the system, the change of pressure
l00rr==========~------------.
r~105O% along the flow path can be calculated as we will see in the section on mechanical
(p"517 to 539 kglm') energy balance.
e r;571060%
(1)''421 10447 kglm') The additional flow invariants - the distance of the plug from the wall and both
.. r=8710 70% flatness measures - are exactly the same as in the incompressible case discussed in
(1"'324 to 355 kglm') the previous chapter.

7.1.3 Volume Equalized Power Law


In most cases the volume equalized power law containing only two parameters is
sufficient to describe the rheology of a given foam:

10 100 r
~= KVE
(y)n
~ (7.7)

Figure 7.2 Volume equalized log-log plot of the rheological data for HPG 40 foams (after [15]) where the flow behavior index, n and the volume equalized consistency index, K VE,
are constants for a given gas-liquid pair at a given temperature [5]. Table 7.1 shows
the relevant variables characterizing the foam flow in tubes where m is the mass flux
For the no-slip condition the velocity profile can be obtained from the integral
and PI is the liquid density.

f [k~ (~ i "v) 1""


Because of the closed analogy between incompressible rheological models and
their volume equalized counterparts, one can readily modify the above relations for
.(y) ~, y' '; - dy', (7.4) other geometries. A similar solution is available for the volume equalized Bingham
plastic model [5].

Note that the volume equalized wall stress is constant and the result of the integra-
Example 7.1 Determination of the Volume Equalized Power Law
tion does not change along the flow path. The velocity profile is linearly "stretched"
with increasing e. Parameters
Let m denote the mass flow rate per unit cross sectional area of the pipe, m ::;:: Enzendorfer (7] measured the pressure drop of N2 foam with base liquid 0.48 mass %
uavgP ::;::const. Proceeding the same way as in Chapter 5, we can derive the usual HPG solution in water at room temperature (21C) and elevated pressures. Figure 7.2
relations between pressure gradient and flow rate. Since neither the pressure gradient
nor the volumetric flow rate is constant along the flow path it is more convenient to Table 7.1 Formulas for tube flow of foams
look for relations containing only flow invariants. The yield stress to wall stress ratio 1: 3)n
(and hence the distance of the plug from the wall) is constant along the flow path: f = 2.+1 (
n KVEPtl-nD-nmn-2

Umax 1+3n
A.::;:: .y,VE' ::;::2ry,vEPI
(7.5)
-=--
' f' m 2 ' u.vg l+n
rw (1 + 2n)(3 + 5n)
OiKE =
and (not surprisingly, if we remember where the principle of volume equalizing 3(1 + 3n)2
comes from) the friction factor has the same property (Valko and Economides [6]):

f = 2n+1ptn K vED-nmn-2(1- )-l

(1+n)(1+2n)(1+3n) ]n
x [ n(l _ )2n2 +
3n + +
1) (2n2 + 2n) + (2n2)2) (7.6)

For a typical flow problem, where the geometry, D, L, the properties of the fluid,
PI,K VE, r y, VE, n, and the mass flux, m, are specified, one arrives at a system of
two equations (Eqs. 7.5 and 7.6) containing two unknowns: f and . Once these
~.---- .. ---------------------------- --------

152 Advanced topics of rheology and fluid mechanics Accounting for mechanical energy 153

shows a log-log plot of the volume equalized wall stress versus volume equalized Reynolds number was calculated from the volume equalized power law according
nominal Newtonian shear rate for his experiments covering the pressure range 3 to to Table 7.1. This calculation method needs no additional parameters except for the
7 MPa, quality range 0.48 to 0.7 and density range 324 to 539 kg/rrr' (Enzendorfer two parameters of the volume equalized power law.
et al. [8]). Determine the volume equalized power law parameters describing the exper-
imental results. 0
Example 7.2 Friction Factor for Turbulent Foam Flow
Solution A water based 0.48 (mass) % Hydroxypropylguar solution is foamed by Nz gas. Deter-
The procedure suggested by Winkler et al. [9] consists of fitting a straight line to the
m
mine the friction factor if the mass flux is = 2000 kg/(m2s), the base liquid density is
log-log data. The flow behavior index is equal to the slope of the line, n = 0.34, and
PI = =
1000 kg/rn", the tube diameter is D 0.055 m, and the volume equalized power
law parameters are n = 0.34 and KVE = 2.5 Pa- s",
=
the pipe consistency index is equal to the intercept, K p. VE 2.S6 Pa . s" . Applying the
relation between the pipe consistency index and the true consistency index, (Eq. 5.56),
we obtain
Solution

From Table 7.1 the (volume equalized) generalized Reynolds number is

23-n on p7-1mn-2
= 22.66 x 0.0550.34 X 1000-0.66 X 20001.66
7.1.4 Turbulent Flow of Foam N
Re
-
>:

K V
(1+3n)"
-n- 2.5 x
(1+3 0.34
X 0.34)34 .
= 1626.

A unified view of the different flow regimes for foams was established by Blauer
et al. [3] who assumed that the rheology of foams obeyed the Bingham plastic
model with yield stress and plastic viscosity depending on quality. According to Since the flow regime is not laminar, the wall Reynolds Dumber has to be calculated:
their method the effective viscosity determined with respect to a given shear rate
can be used to obtain a generalized Reynolds number which, in tum, determines the NRe=
1 + 3n ) NRe- _ (1 + 3 x 0.4) 0.4 1626 _- 9660 .
( -- 4n 0.4
X
friction factor. The turbulent behavior is described by the Newtonian theory. The
method gives higher than realistic friction pressures for foamed polymer solutions
Applying the explicit approximation to the Virk MDRA (Eq. 6.9) the friction factor is
since no drag reduction is taken into account.
The description of the laminar and the turbulent flow is separated in the work
of Reidenbach et al. [4]. Their turbulent flow analysis is based on the extension
x = In[ln(NRe)l = 2.217
of the Melton and Malone [10] procedure, incorporating an additional dependence InU)::: 28.135 + (-29.379 + (S.2405 - 0.86227x)x)x = -5.S9
on density: f = exp( -5.91) = 0.002S.
Dt;.p = A' x De
4L Pf
(8UD avg)m
'
(7.8)
It is interesting to compare the friction factor with the one calculated from the widely
used correlation, Eq. 7.9, which gives
where the exponents were found to satisfy the additional constraints x = m - 1 and
e = m. Rewriting the above correlation into friction factor form we obtain f :::2 x S" x A' x (pu,vg)m-2 = 2 x SI.23 x 0.109 X 2000-0.77 = 0.0081,
f = 2 x 8m x A' x (puavg)m-2, (7.9) i.e., a three times larger value. 0

indicating that the experimental friction factors of reference [4J depend only on
the product (puavg), but not on the individual values of P and u.vg The correlating 7.2 Accounting for Mechanical Energy
parameters for foams of an 0.48 % HPG fluid (rewritten from Table 6 of reference [4]
into SI units) are m = 1.23 and A' = 0.109 [kg/(mz. s)]O.77. 7.2.1 Basic Concepts
Winkler et al. [9] found that (a) non-laminar foam flow can be well described
in the framework of the volume equalized principle and (b) large-scale turbulent Mechanical energy losses per unit mass of flow are determined by the rheology and
flow of HPG foams shows typical drag reduction phenomena. They applied the Virk it is possible to write an accounting balance. The balance along the flow path results
MDRA (in the form of Eq. 6.9) to obtain friction factors. The corresponding wall in the desired relation between end pressures and the mass flow rate.
154 Advanced topics of rheology and fluid mechanics Accounting for mechanical energy 155

The Bernoulli equation (for flow in a circular pipe) can be written as [11] Under these assumptions the integral of the mechanical energy balance from inlet
1 to outlet 2 (being at a distance L) yields [6]
g(z} - Z2) - 4 [U;Vg.1 _ U;Vg.2] _
ctKE.I ctKE.2 I,
r dpP _ l,r 2ju~vg
2
D
dt = 0, (7.10)
(7.15)
where Z, l, U, p, p, D, I, xs are the vertical coordinate, length coordinate, linear
velocity, pressure, density, diameter, friction factor and kinetic energy correction
factor, respectively. As we saw already, the kinetic energy correction factor where
(Uavg)3 1
se == -( 3) , (7.11)
KI = 2jbm2
U avg

characterizes the flatness of the velocity profile and the friction factor is defined in -1
terms of the wall stress by K2=--
LW
21aKE
1= -1-2-' (7.12) 1 a
:zPUavg
K3 == 21ctKE - 2jb2m2'

7.2.2 Incompressible Flow For vertical flow a similar but somewhat more complicated solution is avail-
able [5]. Equation 7.15 can be solved for one of the pressures, if the other pressure
If the fluid is incompressible, then the density is constant and the isothermal mechan- is specified, or for m, if the two pressures are known.
ical energy balance contains only the constant pressure gradient, the potential energy
change and the pressure change. We can use all the rheological relations substituting
the pressure drop, ll.p, by the potential difference, ll.p - pgll.z. Example 7.3 Pressure Drop Calculation for Foam Flow
Determine the outlet pressure, P2, from an L == 100 m long, D = 0.055 m diameter
7.2.3 Foam Flow horizontal tubing, if the inlet pressure is PI = 1 MPa and the inlet How rate of the
foam is q = 0.0125 m3/s. Assume isothermal flow.
Foams are compressible and the neglecting of the kinetic energy change may result The following additional information is available: The base liquid density is PI =
in a considerable error, especially at lower pressures. To simplify the treatment 1000 kg/rrr' The foam density was determined at two pressures: at the inlet pressure
. 3 I
consider isothermal foam flow in a horizontal tube. By virtue of the principle of PI = 190 kglm3 and at another pressure, P3 = 2 MPa, P3 = 320 kg/m . The va ume
volume equalization both the friction factor and the kinetic energy correction factor equalizedpower law parameters are n = 0.4 and K vs = 1.6 Pa . s".
are constant along the flow path and, thus, the following differential form of Eq. 7.10
can be derived:
Uavg dUavg dP 21 U;vg Solution
._!'--~ - - - -- dt = O. (7.13)
ctKE P D The firststep is to describethe volumetricbehaviorof the foam at the given temperature.
Before progressing further one has to describe the volumetric behavior of the foam. We know the foam density at two pressure levels, and hence the two parameters of
Once entering the tube, the mass ratio of the gas to liquid is constant along the flow Eq. 7.14 can be readily determined:
path. The only variable affecting the specific volume of the foam is the pressure.
While the level of description might vary from simple to very sophisticated, one
can assume that the specific volume of the foam along the flow path is described
a= (2. _ 2.)/ (..!.. - ..!..) = 103
PI P3 PI P3
4.32 x m2/s2

sufficiently by the relation


1 a
- = - +b, (7.14)
b = (PI _ P3)
Pi P3
(_1_) = 9.48 10-
PI - P3
x 4rn3/kg.

p P
where a and b are constants. Such relation holds, e.g., if the gas obeys the virial The mass flux is
equation of state (truncated after the second term), the liquid is incompressible and m= qp~ = 1000 kgl(m2 s).
the change in the amount of the dissolved gas can be neglected [5]. D2-
4
156 Advanced topics of rheology and fluid mechanics Rheometry 157

The friction factor and the kinetic correction factor is calculated according to Table 7.1: difference). The flow equations of Chapter 5 are readily applicable. The flow curves

f -- . (1 +
2nTI 3n)"
-n- k VEP!t=tr: m-
.n-'
observed in pipes of different lengths and diameters should collapse into one curve .
For some fluids this does not happen, indicating the violation of one or more of the
main assumptions. The reason for the entrance and wall effects may be different.
= 24T. ! ( 1+ 30.:' 04) 0.4
x 1.6 X (03)1-0.40.55-0.4(103),,-2 = 0.0266 For crosslinked gels and especially for foams the wall effects are more important.
Most likely the different composition of the near wall part of the fluid is responsible
(1 + 2n)(3 + 5n) for the wall effect. Whatever the microscopic reason is, what we see is the shift of
OiKE =
3(1 + 3n)2 = 0.620. the flow curves along the shear rate axis with decreasing diameter of the tube. The
phenomenon can be well treated assuming an apparent slip velocity at the wall. Then
The coefficients in Eq. (7.15) are
the observed flow rate is the sum of two flow rates:
1
Kl = 2fbm2 = 0.0195 [l/Pa] (7.16)
-1 The first part is the one related to shear flow, the second part is due to slip at the
Kz = -- = -30.3
2faKE wall. To put Eq. 7.16 in terms of nominal shear rates we divide by the cross sectional
KIa 4
area and multiply by 8/ D. The resulting equation is
3 = -2f - --,-.-, = -8.75 x 10 .
OiKE 2fb-m-
After substitution, Eq, 7.15 becomes ( 8uD avg)
obs
= (8U avg)
D true
+ (8Us)
D
, (7.17)

106
0.0195 X (l06 - pz) + (-30.3) In- where Us is the slip velocity, i.e., the velocity of the fluid at the wall.
P2 A procedure called slip correction (Cohen [12]) is intended to identify and cut
+ (-8.75 x 104) In 4.32 x 103 +9.48 x 10-4 x 106 = 100 . off the second term on the right hand side and hence reveal the corrected part of the
4.32 x 103 + 9.48x 10-4 x pz 0.055 nominal Newtonian shear rate. To do this some assumptions on the slip velocity are
inescapable.
Th~ solution of the above equation (found by a numerical method) is P2 = 0.974 x
10' Pa. It is readiI y verified that the flow is laminar since the generalized Reynolds
number is
16 7.3.2 Slip Correction
NRe.Vt = f
= 601.
Mooney method [13J
It is interesting to know whether the kinetic correction factor plays any significant
role in the above calculation. The reader can verify that if the kinetic correction factor The key assumption is that the slip velocity depends only on the wall stress;
is taken as unity, the calculated outlet pressure is 60% higher. The discrepancy is about
100%. if the kinetic energy change is totally neglected, i.e., the kinetic energy correction Us = j3T,.... (7.18)
factor is taken as a practically infinite number in the above calculations.
Correctly accounting for the kinetic energy change decreases the calculated outlet where j3 is the slip coefficient (which itself might depend slightly on the wall stress).
pressure. Indeed, the quality at the inlet is about 82% and it increases to approximately This is a reasonable assumption for fluids without macroscopically observable struc-
98% at the outlet. Consequently, the average velocity at the outlet is much higher than tures. Rewriting Eq. 7.17 in terms of the slip coefficient leads to
at the inlet (approximately 10 times) and the increase of the kinetic energy is manifested
by additional pressure loss. 0
(
~-
8uavg) - (8U
_ ~-avg) + 8j3r,,-. 1 (7.19)
D obs D true D
7.3 Rheometry The plot of the observed nominal Newtonian shear rate versus reciprocal diameter
of the pipe at a fixed wall stress would yield a straight line. The slope divided by
7.3. 1 Pipe Viscometry 8Tw is the slip coefficient, j3 (characteristic for the given wall stress.) Note that for
the given wall stress this is an extrapolation to the infinite diameter pipe (where the
The straightforward way to measure rheological properties is to establish tube flow distortion caused by slip is negligible) and the optimal estimate of the true shear
and measure both the volumetric flow rate and the pressure drop (or potential rate is nothing else but the intercept. In practice the intercept is not used; instead the
~--."- .... ----------------------------- -----------------._--", ..__ ._--

158 Advanced topics of rheology and fluid mechanics Rheometry 159

same procedure is repeated for other wall stresses, and a suitable smooth curve is Table 7.2 Observed rheological data [7]
drawn through the slip coefficient versus wall stress points. The curve then gives a
D [em]
slip coefficient for any wall stress. Once the slip coefficient is known as a function
of the wall stress, the original measurements can be corrected by subtracting the 1.2 0.95 0.8 0.6 OA05
second term in Eq, 7.19. . u/d 1"w u/d 1"", u/d 1",. uJd rw uJd rw
Technically, slip correction should be made only if pipes of at least three different [I/SJ [Pal [lIS] [pa] [lIS] [pa] [lIS] [pa] [IISJ [PaJ
diameters are utilized and the flow curves cover (at least partly) the same wall stress 9.5 6.3 19.2 8.6 32.1 11.0 76.1 14.4 247.4 18.6
interval. Since usually the measurement points do not correspond exactly to the 19.0 10.2 38.3 13.3 64.2 15.5 152.2 19.8 494.7 26.1
same wall stress, interpolation may be necessary. Possible negative intercepts, large 28.5 13.2 57.5 15.7 96.3 18.8 228.2 23.4 742.1 32.6
38.0 15 76.7 17.1 128.4 20.8 304.3 26.1 989.5 37.7
scatter around the straight line and hectic variation of the slip coefficient with the
47.5 16.8 95.8 19.1 160.5 22.8 380.4 28.8 1236.8 42.5
wall stress may indicate that the data do not support the basic assumption concerning 57.1 17.7 115.0 20.6 192.6 24.4 456.5 30.9 1484.2 45.6
the dependence of the slip velocity on the wall stress.

Example 7.4 Slip Correction of Foam Rheological Data


Oldroyd-Jastrzebski method [14J
Enzendorfer [7] measured the pressure drop of Nz foam with base liquid 0.48 mass %
For fluids with macroscopic structure such as slurries or foams the apparent slip is HPG solution in water at room temperature (21 DC) and elevated pressures. Consider
the result of a more complex interaction between the wall and the fluid and, hence, his measurements (Table 7.2) at pressure p = 5 MPa and quality I" 0.585. (Since the =
also the diameter of the pipe itself affects the apparent slip. The basic assumption is =
foam density is p 447 kg/nr' and the liquid density is 1000 kg/m", the corresponding
that the slip velocity depends on the wall stress and on the diameter according to specific volume expansion ratio is e = 1000/447 = 2.24.)
Determine whether (apparent) slip at the wall should be taken into account. If the
answer is yes, decide which slip correction method is suitable. Determine the slip model
(3.20) and correct the measurements for slip. Plot the corrected flow curve.

where f3c is the modified slip coefficient (possibly slightly depending further on the Solution
wall stress.) Rewriting Eq. 7.17 in terms of the modified slip coefficient leads to
The observed wall stress versus nominal Newtonian shear rate curves are shown in
Figure 7.3. The points corresponding to different diameters do not lie on one curve,
+ B,BcTw~. indicating the presence of a wall effect. The nominal shear rate corresponding to a
( BUavg) = (Buavg) (7.21)
given wall stress is higher in smaller pipes, which is typical for the (apparent) slip
D obs D true D
phenomenon.
According to the above relation a plot of the observed nominal Newtonian shear
rate at a fixed wall stress versus the reciprocal diameter of the pipe squared yields 50
a straight line. The slope divided by BTw is the modified slip coefficient, f3c (charac-
40
V 1-0
teristic for the given wail stress).
Again, this is an extrapolation to the infinite diameter pipe (where the distortion ./>'
V
o [em]
caused by slip is negligible) and the optimal estimate of the true shear rate is nothing
J' V
V --0--1.2
else but the intercept. In practice the intercept is not used; instead the same procedure
is repeated for other wall stresses. It is anticipated that the wall stress, the modified
10
fl~ -<>-0.95
---0-0.8
-<>-0.6
slip coefficient should be constant or varying linearly. Once a suitable straight line
-<>-0.405
is determined the modified slip coefficient for any wall stress can be computed and
o
the original measurements can be corrected according to Eq. 7.21- o 200 400 600 800 1000120014001600
For slurries, the Oldroyd-Jastrzebski slip correction has been found the only aulD [s-11
suitable method (Hanks [15]). According to recent evidence the same method may Figure 7.3 Wall stress versus observed nominal Newtonian flow rate in different pipes
be used to correct for the apparent slip of foams (Enzendorfer et al. [8]). (after [7])
160 Advanced topics of rheology and fluid mechanics Rheometry 161

Selecting different wall stress levels (from 12 Pa to 26 Pa) we can read off the 4Xla5rr==================~--------~
nominal shear rates corresponding to different pipe diameters. The points are plotted
Pc= a +b.w
versus the reciprocal pipe diameter in Figure 7.4. From this Mooney plot we can
a = 8.6 10-6 m-2/(Pa.s)]
see that the optimal estimate of the shear rate in an infinitely large pipe (i.e., the
intercept of the straight lines with the vertical axis) is negative for virtually all wall b = 7.010-7 m-2/(pas)]
stresses involved. Moreover, there is an inversion phenomenon: The optimal estimates
of the true shear rates increase with decreasing stress, which is physically impos-
sible. It is obvious that the data do not support the slip model on which Eq. 7.18
is based.
The same data plotted with the reciprocal diameter squared as the abscissa are
shown in Figure 7.5. The Oldroyd-Jastrzebski method gives consistent estimates for

O~---- __~ ~ __-L __~ ~-~


700
o 10 20 30
600 ;.(Pal
TW [Pal
0 30
0 28
500 C. 26
Figure 7.6 Modified slip coefficient vs. wall stress for Example 7.4 [8]
V 24

'I
~
400
..
0 22
20

Cl 300 .,
X 17
15
the infinitely large diameter [8]. The slopes of the straight lines divided by 8t',.. are the
'<,
::>
co
- 13 modified slip coefficients, fic. A subsequent plot of the modified slip coefficient versus
200 the wall stress (Figure 7.6) shows a linear trend. The equation of the straight line is
100 used to correct the measured nominal shear rates:

0
(
8Uavg)
(
8uav s) _ (~;) [a+brw]
D corrected D observed
50 100 150 200 250
l/D [m-1] a = 8_6 x 10-6 __m2
Pas
Figure 7.4 Mooney plot of the data (after [71)
b = 7 X 10-7__m2
Pas
700 The corrected data points are shown in Figure 7.7. The points collapse into one curve
",!pal which is the "true" pipe flow curve of the studied foam [8]. (Note that the corrected
600 0 30
0 28 data points, together with other measurements, were used in Example 7.1 to derive the
C. 26
v
volume equalized power law parameters.) 0
500
0
+
'I The most widespread instrument to measure fluid flow properties is the rotating-
~ 400 .,
X
cup viscometer. The evaluation of the data is more difficult than in the case of
Cl
.......
:;) pipe viscometry. The rea] shear rate at the bob cannot be revealed by first order
CO
derivation of the experimental data. A Rabinowitch-Mooney type analysis would
require higher-order derivatives. In addition, the presence or absence of slip is
more hidden than in the case of pipe flow curves. These problems may be over-

0~~~=3.
o 10000 20000 30000 40000 50000 60000 70000
come partly by minimizing the gap between the bob and cup, but in the case
of slurries and foams we are limited by an additional constraint: The gap size
should be significantly larger than the characteristic size of the particle or bubble.
I/D2 [m-2] In view of the difficulties above it is not surprising that the application of rota-
Figure 7.5 Oldroyd-Jastrzebski plot of the data for Example 7.4 [8J tional viscometry in hydraulic fracturing is restricted at most for routine quality
162 Advanced topics of rheology and fluid mechanics References 163

11. Denn M.M.: Process Fluid Mechanics, Prentice Hall, EnglewoodCliffs, N.J., 1980.
~r.=====~----~-----------,
o D_1.2 ....
12. Cohen, Y.: Apparent Slip Flow of Polymer Solutions, in Encyclopedia of Fluid
o 0 .15 CJrI Mechanics,Cheremisinoff,N.P. (ed.) Vol. 7, Gulf, Houston, TX, 1986.
~ D_.8 em
V 0 .8 .... 13. Mooney, M.: Explicit Formulasfor Slip and Fluidity,J. Rheology, 2(2), 210-222, 1931.
<> D_.405_ 14. Jastrzebski,Z.D.: Entrance Effects and Wall Effects in an Extrusion RheometerDuring
the Flow of ConcentratedSuspensions, Ind. Eng. Chem. Fund., 6 (3), 445-453, 1967.
15. Hanks, R.W.: Principles of Slurry Pipeline Hydraulics, in. Encyclopedia of Fluid
Mechanics, Cheremisinoff,N.P. (ed.), Vol. 5, 237-240, Gulf, Houston, TX, 1986.

(8u/D)trw Is-11
Figure 7.7 Correctedflowcurve for Example7.4 [8]

control. Fundamental rheological characterization should rather rely on pipe flow


measurements.

References

1. Cameron,1.R. and R.K. Prud'homme, Fracturing-Fluid Flow Behavior, in Recent


Advances in Hydraulic Fracturing, SPE MonographSeries, Gilded, J.L, Holditch, S.A.,
Nierode, D.E and Veatch,R.W. Jr. (eds), SPE RichardsonTX, 177-209, 1989.
2. Assar G.R., Burley R.W.: Hydrodynamicsof Foam Flow in Pipes, Capillary Tubes, and
Porous Media, in: Encyclopedia of Fluid Mechanics, Cheremisinoff,N.P. (ed.) Vol.3,
26-42, Gulf, Houston,TX, 1986.
3. Blauer, R.E., Mitchell, B.l. and Kohlhaas, C.A.: Determinationof Laminar, Turbulent,
and TransitionalFoam Flow Losses in Pipes, Paper SPE 4885, 1974.
4. Reidenbach,V.G., Harris, P.C, Lee, Y.N. and Lord, D.L.: Rheological Study of Foam
FracturingFluids Using Nitrogen and CarbonDioxide, SPE Production Engineering, 1,
39-41, 1986.
5. Valk6 P. and Economides M.J.: Volume Equalized ConstitutiveEquations for Foamed
Polymer Solutions, J. Rheology, 36, 1033-55, 1992.
6. Valk6 P. and EconomidesMJ.: Accounting for mechanical energy in steady-state
laminar foam flow, Proc. 4th Eur. Rheol. Cant, Sevilla 49, Sept., 1994.
7. Enzendorfer,C.: Foam Rheology,MS Thesis, IDP, MiningUniversity Leoben, Leoben,
Austria, 1994.
8. Enzendorfer,C., Harris, R.A, Valko, P., Economides,M.J., Fokker, P.A., Davies, D.O.:
Pipe viscometry of foams,J. Rheology, 39 (2),345-358, 1995.
9. Winkler,W., Valk6,P. and Economides,MJ.: Laminar and Drag Reduced Polymeric
Foam Flow, J. Rheology, 38, 111-127, 1994.
10. Melton, L.L. and Malone,W.T.: Fluid MechanicsResearchand EngineeringApplication
in non-NewtonianFluid System, SPEl, (March), 56,1964 .
--" ..-~--"-"'-~'--"----"-----="'-----"-'-'-----

8
MATERIAL BALANCE

8.1 The Conservation of Mass and its Relation to


Fracture Dimensions
In creating a hydraulic fracture, large volumes of fracturing fluids are injected. These
fluids are expected to generate an appropriate fracture volume while a portion of the
fluids is lost (or "leaks off') into the porous medium through the faces of the fracture
that is being created.
In practice, fracturing fluids are injected in various stages, differing from each
other in terms of chemical composition (and thus, rheological and leak-off properties)
and/or functionality. In propped hydraulic fracturing, some of the fluid stages are used
as proppant carriers.
In this chapter a material balance is employed describing the link between the
injected volume, the created fracture. volume at the end of pumping and the leak-off
volume. Mechanisms intended for the description of leak-off are presented in detail.
While material balance itself is not intended as a predictive or modeling tool, it
is an essential element of any fracture propagation model. In addition it provides a
means of consistency check for modeling results even if the details of the calculations
are not available. As such, the concepts of this chapter are critical to the understanding
of the actual models presented in the subsequent chapters.
An overall material balance for a hydraulically induced fracture can be written as

(8.1)

stating that the volume of the injected fluid, Vi, equals the sum of the created fracture
volume V, and the volume of fluid, V L, "lost" across the fracture faces into the
formation. The latter is often termed as the leak-off volume. Equation 8.1 neglects
the compressibility of the fluid and does not make a distinction between the volume
of the fracture and the volume of the fluid contained within, because secondary
effects (such as a possible small unwetted zone at the tip) have little significance in
material balance calculations considered in this chapter.
In practice, the injected volume is that pumped down the well and Eq. 8.1 would
imply that the created fracture volume consists of the volumes of two individual
fracture wings. For modeling purposes, however, it is more convenient to deal with
166 Material balance The conservation of mass 167

only one of the wings, especially because the distance from the well to fracture tip and
is conventionally selected as the key variable. Therefore, in this book the variables Aw
T}=-.-, (8.5)
refer to one fracture wing of a two-wing fracture. In particular, Vi is the injected It
volume into one fracture wing, V L is the volume of fluid entering the formation respectively. .
through the two created fracture surfaces of one wing and V is the volume of fluid In many cases there is evidence that the created fracture remains in a well defined
contained in one fracture wing. These variables correspond to a given instant in time. lithological layer, and the fracture is characterized by a constant height, hf. In this
The strongly related characteristics of a hydraulically induced fracture are the case the length of the fracture, xf, is simply
fluid efficiency, the fracture surface and the average width. The efficiency, T/, is the
fraction representing that part of the fluid remaining in the fracture: A
xf=- (8.6)
V hf
T/= Vi' (8.2)
The words fracture half-length and fracture length are used interchangeably. If a
and is between zero and unity (or between zero and one hundred, if expressed in constant hf cannot be assumed, the fracture length is the largest distance between
percent). The fracture surface, A, is half of the surface area of the fluid body of the well and a point on the fracture tip.
one wing, i.e. it is the area of one face of one wing. From the point of view of the It is also convenient to make a clear distinction between the values of the above
surface area, the fracture wing is envisioned as a body limited by two large parallel variables at any time, t, and at the end of pumping, i.e. at time te' We will use the
faces. Since the width is small relative to the other dimensions, we do not make any subscript e to emphasize that a given value corresponds to teo
distinction between the summed area of the faces and the total area. The average The above variables and relations constitute an essential and basic common
width, w, is defined by the relation language with which hydraulic fracturing workers should communicate.
As in other disciplines, basic unassailable principles should override the descrip-
V=Aw. (83) tion of individual phenomena. Warpinski et al. [1] noted that "In recent years, there
has been a proliferation of fracturing simulators used in the oil industry. This prolif-
With Eqs. 8.2 and 8.3 the efficiency, fracture surface and average width depend
eration was intensified by the availability of personal computers and the need for
on each other if the injected volume is specified.
fast running design simulators for use in the field. Applying these models as 'black
To remain consistent, we denote by i the injection rate entering one of the wings
boxes', without knowing the underlying assumptions, may lead to erroneous conclu-
(Figure 8.1). Very often the injection rate, i, is constant and hence, the injected
sions, especially for the unconfined fracture growth."
volume up to a given injection time, t, is given by Vi = it. (Usually the total injection
Equations 8.1 to 8.6, though seemingly trivial, may provide a useful means to filter
rate, 2i, is specified in engineering practice.) In the constant injection rate case,
out major inconsistencies of computed results, even if the details of the calculations
instead of Eqs, 8.1 and 8.2 we can use
are not known. To illustrate this point, we will use the comparison of hydraulic-
(8.4) fracture models presented in the above cited study by Warpinski et al. [1].

Example 8.1 Simple Consistency Check of Model Results

AppendixA contains details and results of the test problem of the Warpinski et al. [1]
comparisonstudy. Use the combination of Eqs. 8.5 and 8.6 to check the consistency
of the results providedby the modelers.
Tables Al and A.2 show the input data. Table A3 contains the results of 34
differentsets of results provided by participants.The results of Table A3 correspondto
fixed height models.According to the problem specification(in addition to others) the
following data were given: injected volume, Vi = it. = 794.9 m3, and fracture height,
hf = 51.8 m. (For unexplainedreasons, some of the participants did not use the value
of hf = 51.8 m, but assumed another height. Use the height provided by the modeler
Figure S.l Schematicof fracturedimensionsandinjectionrate
distribution in the consistencycheck.)
---------=--------_.---._------_._---_.

168 Material balance Fluid leakoff and spurt loss as material properties 169

Solution where subscript c denotes a calculated value. The calculated efficiency is shown in the
last column of Table 8.1.
In Table 8.1, we present the x], IVand 1) values provided by the modelers, distinguishing Some discrepancy between the modeler-provided and the calculated values may
them by the subscript m. Of course, these three quantities are not independent, once V; be attributed to the finite number of digits. A larger discrepancy means, however,
and h , are specified. that certain modelers either do not apply rudimentary criteria or do not use similar
The combination of Eqs. 8.5 and 8.6 yields the following expression for the effi- terminology. For example, the deviation in Case 1 is difficult to explain. An efficiency
ciency: greater than 100% as in Case 3 indicates that a fracture with volume larger than the
hf.mXfmWm injected volume was created. In Cases 12 and 13 a fracture only half as large as that
1)e = indicated by the modeler's efficiency was created. The above results underline the
V;
danger of relying on a computer code without cross checking as Warpinski et at. [1J
pointed out. 0
Table 8.1 Characteristics provided by the modelers (m) and
calculated efficiency (1),.)

hj.m xJ~m Wm 11m 11e 8.2 Fluid Leakoff and Spurt Loss as Material Properties
m m m % %
The power of material balance can be realized only if we are able to describe the term
51.8 774.8 0.0154 85.5 77.6
VL. There are two main schools of thought concerning leakoff. The first considers the
2 51.8 1480 0.0073 72.3 70.8
3 62.2 787.6 0.0185 93.0 114.2 phenomenon as a material property of the fluid/rock system. The second considers
4 51.8 810.5 0.0157 83.1 83.2 leakoff as a consequence of flow mechanisms into the porous medium and uses a
5 51.8 1374 0.0081 72.2 72.8 corresponding mathematical description.
6 51.8 697.4 0.0188 85.4 85.4
7 51.8 1159 0.0102 76.6 76.8
8 51.8 830.3 0.0155 84.0 83.9 B.2.1 Carter Equation I
9 51.8 1231. 0.0094 75.0 75.4
10' 61.0 755.9 86.0 The polymer content of the fracturing fluid is partly intended to impede the loss
11' 61.0 1267. 77.0 of fluid. The phenomenon is envisioned as a continuous build-up of a thin layer
12 51.8 410.6 0.0152 81.9 40.8 (the filter cake) which manifests an ever increasing resistance to flow through the
13 51.8 618.4 0.0091 73.0 36.9 fracture face. In reality, the actual leakoff is determined by a coupled system, of
14 51.8 1401. 0.0081 73.8 74.2 which the filter-cake is only one element. (The other elements will be considered
15 51.8 674.2 0.0196 85.9 86.0
in more detail in the next section.) A fruitful formalism dating back to Howard and
16 51.8 827.8 0.0152 82.5 82.2
17 51.8 0.0094 74.4 74.4 Fast [2] is to consider the combined effect of the different phenomena as a material
1215.
18 51.8 1178. 0.0098 75.0 75.5 property. According to this concept, the leakoff velocity, VL, is given by the Carter
19 51.8 1084. 0.0110 75.0 77.9 equation (See Howard and Fast [2]):
20 51.8 774.8 0.0152 6L8 77.0
21 51.8 1411. 0.0071 73.6 65.4 (8.7)
22 62.2 766.9 0.0190 93.0 114.3
23 51.8 639.5 0.0208 86.4 86.8
24 51.8 1255. 0.0091 74.3 74.8 where CL is the leakoff coefficient, measured in units of m/sl/2 and t is the time
25 51.8 551.1 0.0246 88.3 88.5 elapsed since the start of the leakoff process. The Carter equation is postulated from
26 51.8 1035. 0.0117 79.0 78.8 empirical observations as shown in Example 8.2. The integrated form of Eq. 8.7 is
27 51.8 652.9 0.0206 89.0 87.6
28 51.8 1020. 0.0119 79.0 79.4 VL ..Jit+Sp,
29 51.8 1233. 0.0097 76.9 77.6 -=2C L (8.8)
AL
30 51.8 619.0 0.0213 86.0 86.1
31 51.8 702.3 0.0186 85.2 85.2 where V L is the fluid volume that passes through the surface AL during the time
32 5L8 1114. 0.0105 76.5 76.6 period from time zero to time t. The integration constant, S p, is called the spurt loss
33 51.8 1035. 0.0116 78.0 78.1 coefficient and is measured in meters. It can be considered as the width of the fluid
34 51.8 961.6 0.0128 81.7 80.2
body passing through the surface instantaneously at the very beginning of the leaksff
'The necessaryinformationhas not beenprovidedby the modeler process. The two coefficients, Cv and Sp, can be determined from laboratory tests.
~---' _----------
..

170 Material balance Fluid leakoff and spurt loss as material properties 171

Example 8.2 Determination of the Leakoff Parameters from the slope, m, and the intercept, b, of the straight line are m = 6.9 x 10-5 m/sl/2 and
Laboratory Data b = 2.4 X 10-3 m.
Comparing Eq. 8.8 with the obtained m and b shows that
Fracturing fluid, pressurized to a representative pressure, flows through a core sample m
with cross sectional area of 20 cm2. The filtrate (loss of fluid) is recorded during one CL = - = 3.5 X 10-5 rn,/SI/2(= 8.8 X 10-4 ft/rninlf2) and
2 .
hour as shown in Table 8.2. Determine the two parameters of Eq. 8.8.
Sp = b = 2.4 X 10-3 m(= 0.1 in.).

Solution A closer look at Figure 8.2 reveals that the spurt loss is a matter of convenience
rather than a physical reality. It is the consequence of our Willingness to describe the
First we compute the square root of time for every point and plot the filtrate volume points by a straight line and our indifference toward the finer details of what actually
divided by the cross sectional area, VLIAL, VS. the square root of time. From Figure 8.2, happens at early time. 0

An additional note concerning the time t is necessary here. Equations 8.7 and 8.8
Table 8.2 Measurements of
filtrate volume for
mean that the given surface element "remembers" when it has been opened to fluid
Example 8.2 loss and has its own "zero" time which might be different from location to location
on a fracture surface.
miu
8.2.2 Formal Material Balance. The Opening Time
1 5.2 Distribution Factor
2 6.7
4 7.3 A hydraulic fracturing operation may last from tens of minutes up to several hours.
10 8.6 Points of the fracture face near to the well are opened at the beginning of pumping
20 9.7 while the points at the fracture tip are "younger". Application of Eq. 8.8 or of its
30 10.6
differential form, 8.7, necessitates the tracking of the opening-time of the different
40 11.4
50 12.5 fracture face elements. If only the overall material balance is considered, it is natural
60 13.2 to rewrite Eq. 8.1 using the formalism of Eq, 8.8:
(8.9)
E 0.007 where the variable K is the opening-time distribution factor. It is defined by the
-.I
<: material balance itself, i.e.
"':::.J 0.006 (8.10)
>
tV
o 0.005
~ and its value depends on the history of the evolution of the fracture surface, or rather
..
:::J
tn
'2
0.004 on the distribution of the opening-time. In particular, if (1) the spurt loss coefficient is
zero, and (2) all surfaces are opened at the beginning of injection, then K is exactly 2.
...
:::J
Q) 0.003 y = 0.0024 + 0.000069x For the no-spurt-loss case this is an absolute maximum of the factor.
Co The relation of the K factor, efficiency and average width is obtained from Eqs. 8.2,
Q)
E 0.002 8.3 and 8.9:
:::J
(5 K= (8.11)
> 0.001
'0
0 Another form of the same relation,
...J
0
0 10 20 30 40 50 60 w
"1/= , (8.12)
Square root time, tl12 (Sll2) 2KCL'.ji +w
Figure 8.2 Filtrate volume through a core (Example 8.2). Slope provides the leakoff coef- shows that the term 2KCL.fi can be considered as the "leakoff width". The K factor
ficient, intercept provides the spurt-loss coefficient plays an important role in some design procedures discussed in Chapter 9.
172 Material balance The constant width approximation (Carter equation /I) 173

8.3 The Constant Width Approximation Example 8.3 Fracture Surface Calculation for Height Growth Models
(Carter Equation II) (Consistency Check II)
In Example 8.1 we investigated the solutions to a specific design problem where the
The opening-time is denoted by T and every surface element has its own T. If the height was specified. Table A.4 of the Appendix gives the results for variable height
actual time is denoted by t, the leakoff flow rate, corresponding to the given surface models (using three different layers). Estimate the fracture surface from Eq. 8.5 (Ad)
element is and from the Carter equation II (Acz) assuming for the latter that the average width
(8.13) (given by the modeler) is constant during the whole injection period. Compare the two
=
results using the following data: i 0.0662 m2/s (25 bpm per wing), I, 12000 s =
(200 min) and CL = 9.84 X 10-6 m/slJ2(0.00025 ft/minl/2).
Assume .that we know the history of the fracture surface growth, i.e. the function
A(r) and/or its inverse function, rCA). Then the leakoff flow rate through the two
fracture faces is the summation of the different flow rates along the surface elements Solution
of different age:
From Eq. 8.5 the first estimate of the fracture surface is simply

2
1 o
A(I)

.Jt -
CL
rCA)
dA-2
-
l' 0 ~
-- CL (dA)
-
dr
dr
'
(8.14)

(where the factor two now comes from the two fracture faces.) Not all the fluid whatever the shape of the fracture face is. From Eqs. 8.18 and 8.19
injected leaks off and hence the fracture grows. The growth rate of the volume is

dA dw wmi
Ac2 = -,- (z
exp({3) erfc(fJ) + 2f3]
r;; - 1 where f3 = 2C .,fiit; ,
L

w-+A~ (8.15) 4CL~ vn Wm


dt dr '
and the creation of new surface brings about an additional loss due to spurt loss and the actual shape of the fracture face is not relevant. The results are given in
Table 8.3.
. dA
2Sp-. (8.16)
dt Table 8.3 Fracture Surface Estimation from
the Data of Table A 4
Carter [2J formulated the material balance in terms offlow rates. He argued that if
Act , Ad Ad
at time t the injection rate entering one wing of the fracture is i, it should be equal m- m2 Ac2
to the sum of the different leakoff rates plus the growth rate of the fracture volume.
Hence, 1 80.3 75.0 1.07
. . l'
1=2
CL dA
r.-=-dr+(w+2Sp)-+A-.
o .,; t - r dt
dA
dt
dw
dt
(8.17)
2
3
82.6
69.5
85.1
65.7
0.97
1.06
4 46.9 41.4 1.13
To obtain an analytical solution for the constant injection rate case, Carter solved a 5 75.9 76.8 0.99
6 68.4 68.5 LOO
simplified version of the material balance, neglecting the fact that the width increases 0.99
7 83.9 85.1
during the fracture growth. If w is constant during the entire pumping period, i.e. the 8'
fracture has its final width already in the first instant of injection, the solution is [2]: 9 53.8 85.1 0.63
10 78.3 73.3 1.07
A(t) = (w + 2s
2 p
)i [
exp(~2) erfc (~) + 2~ ]
r;; - 1 . (8.18) 11 81.9 82.8 0.99
4CL~ .,;n 12 39.6 51.7 0.77
13 45.5 40.4 1.13
where 14 71.1 71.6 0.99

~= 2CL.Jm
w+2S
. p
(8.19) 15
16
60.2
83.1
60.6
82.8
0.99
1.00
17 70.0 95.2 0.74
Equation 8.18 gives the fracture surface if both the width and the time are specified. 'The necessary information has not been provided
It will be used extensively in Chapter 9. by the modeler
. .0-.__

174 Material balance The power law approximation to surface growth 175

If the two different ways to estimate the fracture surface yield the same result, i.e. where the subscript e refers to the end of pumping. The reader may wonder why we
the quotient is 1, there is no question of consistency(howevera "too good" agreement introduce a new symbol, go(a). Isn't it exactly the opening-time distribution factor k
may indicate that the given model effectively used the Carter equation II even if it at the end of pumping? The answer is yes, but with some restrictions. The function
claimed to be a real pseudo-Sf)or 3D model.) A quotient larger than unity indicates
go(O') can be determined by an exact mathematical method because it involves the
that the leakoff was less than reasonable. 0
assumptions that (1) the surface grows according to the power law Eq. 8.20, (2) the
fluid leaks off according to Carter equation I and (3) the spurt loss is zero.
8.4 The Power Law Approximation to Surface Growth In order to derive a closed form of the function go(O'), consider an elementary
surface, cIA, which is opened at time r. The volume of fluid lost through the elemen-
If we plot the fracture surface computed from the Carter equation II vs. time using tary surface since its opening until time t, is given by Eq. 8.13:
log-log coordinates, the result is always similar to the one shown in Figure 8.3. At
!. t, CL
early times the slope of the curve is unity, and at later times it decreases to
Probably motivated by this fact, Nolte [3] postulated a basic assumption leading to
a remarkably simple form of the material balance. He also considered the constant
dV L = cIA
i
r
;;--:;
...;t - 1:
dt. (8.22)

injection rate case and assumed that the fracture surface evolves according to a While the fracture surface increases from zero to Ae, the volume of fluid leaking off
power law, is the integration of Eq. 8.22 with respect to the surface

(8.20) A'i"
with exponent a being constant during the injection period. In other words he
V Le = 2
l
o r
;;--:;
CL drda,
...;t-1:
(8.23)

assumed a particular form of the solution of the mathematical model which is still Substituting Eq. 8.23 into Eq. 8.21 yields
not even specified, because the actual fracture surface evolution is determined (along
with the material balance) by additional phenomena such as elasticity and fluid flow, 1
go(O') = Ae Jo
r ..jt;1 (it'r.Jt 1
_ 1: dt
)
cIA, (8.24)
as we will see in the following chapters.
where the opening-time is given (by virtue of Eq. 8.20) as
8.4.1 The Consequences of the Power Law Assumption
(8.25)
Nolte [3,4] also introduced a new function

It is convenient to consider dimensionless variables:


(8.21)
(8.26)

From Eqs. 8.25 and 8.26


(8.27)

Substituting Eqs. 8.26 and 8.27 into Eq. 8.24 gives

(8.28)

The integral on the right-hand side can be given in closed form as


Injection time
O'.Jif(O')
(8.29)
Figure 8.3 Log-logplot of fracturesurfacecalculatedfromthe CarterequationII vs. time go(a) = f(~ + a) ,
----- ------ --- ----- ------------. -_. __ ...
_._ .. ------------------------

176 Material balance The power law approximation to surface growth 177

If the first two assumptions of this section are accepted (i.e. power law surface
growth and Carter I Ieakoff), and the exponent a is assumed to be known, the material
balance for any time instant, during injection, can be written in the form

(8.32)

which becomes
0.6 (8.33)
0.4
0.2 at the end of the injection.
The exponent a has been expIicity related to the theological behavior of the
0.5 1.5 fluid by many authors (see Section 9.6). Therefore, it is often considered known
a in design calculations. In spite of this, to our knowledge Eq. 8.32 has never been
used for design purposes, probably because the analytical form (Eq. 8.30) has not
Figure 8.4 The plot of the go function
been known. Instead, an interpolation technique described in Section 8.4.2 has been
preferred in the literature.
where fCO') is the Euler gamma function. Figure 8.4 shows the plot of the function
goCa). Not surprisingly, the function is two when a = 0, i.e. go(a) reaches the abso-
lute maximum of K. Indeed, if the exponent is zero, the whole fracture surface has Example 8.4 Consistency Check In
to be opened at the start of injection and maximum fluid volume is leaked off.
Notice that because of the special properties of the Gamma function (Abramowitz Consider the five-layer results of the comparative study, given in Table A.5 of the
and Stegun [5]), the form Appendix. Compute the opening-time distribution factor, k. Assuming that the fracture
surface has evolved according to a power law, estimate the exponent, <X. Comment on
22aa[fCa)]2 the results from the point of view of consistency.
go(a) = (1 + 2a)f(2a)' (8.30)

given by Valko and Economides [6] is completely identical to Eq. 8.29. Solution
There are two remarkable facts concerning the above result. First, go(Ci) is really As in Example 8.2, first we estimate the fracture surface from Eq. 8.5,
a function of the exponent only. Second,

go(~) = ~ == 1.57 and go(l) = ~ == 1.33, (8.31)


Whatever the shape of the fracture face is, the factor K can be computed from (see
indicating that for two extremely different surface growth histories (Ci = 0.5 and a =
Eq.8.10)
1), the opening-time distribution factor differs by less than 20%. (1 - Tlm)ite
The values given in Eq. 8.31 are often referred to as lower and upper bounds, J(~ = .
2AcCLA
respectively. The reason is that Nolte [3J, and many others impressed by his
pioneering work, postulate that a and TJ are strongly related and that for a dominant (The same value could be obtained directly from Eq. 8.11.) Finally, a suitable expo-
fluid loss (TJ == 0) the value of a will be approximately ~ and for a negligible fluid loss nent, a., is determined solving the nonlinear equation
(TJ == 0) the exponent a will approach 1. It is difficult to argue about assumptions on
the value of an exponent the existence of which is also an assumption. Nevertheless,
we note that once we assume that the fracture surface grows with constant exponent,
we cannot relate this exponent to the fluid efficiency, because the fluid efficiency is for the unknown o.The solution can be obtained by any numerical root-finding method.
not constant during the fracture propagation. It always starts from unity at time zero The results are shown in Table 8.4.
and decreases afterwards. For very long pumping times, the fluid efficiency always Having anticipated the exponent between half and unity, the diversity of the calcu-
approaches zero. lated exponents raises some serious questions. As was noted in [6], the results do not
._- ----_ ... --._ "'--_ .. '---"'---'-.'--. __ .

178 Material balance


Numerical material balance 179

Table 8.4 Opening-time distribution


and (2) that K varies linearly between these bound with the final 7],the global material
factor and estimated power
law exponent for the data balance is often used in the form
of Table A.4
it r: 8
Kc etc - = w + Ccv t[37] + (1 - 1);1"]. (8.36)
A
1 0.91 3.06
2 1.52 0.59 We note again that it is difficult to accept these assumptions because the exponent
3 1.03 2.17 is considered to be constant during propagation while the fluid efficiency decreases
4 0.24 55.22 monotonically. Indeed, how does the fracture surface know which exponent a it has
5 1.38 0.88 to grow with, if this exponent is determined by the final value of fluid efficiency,
6 1.39 0.86
7
which depends very much on the time when the injection is stopped? Nevertheless,
1.33 1.01
S" Eq. 8.36 is used extensively in design calculations (Meng [7]). It is a useful equation
9 3.06 -0.48 because in "design mode", when the fracture surface is specified and the width is
10 0.81 4.00 calculated from principles beyond the material balance, it yields a simple quadratic
11 1.26 1.20 equation from which the necessary time of injection (first its square root) can be
12 1.06 2.05 easily obtained. Since a suitable value of the fluid efficiency is involved, Eq. 8.35
13 0.59 8.41 is used simultaneously with Eq. 8.5 embedded into an iteration loop. The procedure
14 1.56 0.52
15 will be illustrated in Section 9.2.3.
1.41 0.81
16'
17"
18' 8.5 Numerical Material Balance
19"
20 2.25 -0.17 The simple result of Eq. 8.18 should hold exactly if the leakoff obeyed the Carter
'The necessary informationhas no! been equation I and the width opened into its final value at the very first moment. Since
providedby the modeler the latter cannot be true we may wish to investigate the question: Can Eq. 8.18 still
work as a bound on the surface growth?
mean that any of the computer codes are wrong. Since the fracture surface does not grow To answer this question, we develop a simple numerical procedure. We assume
~ccording to a power law in these numerical models, some discrepancy from the antic- that the injection rate is constant and discretize the time with constant time step, b.t.
~pated.interva.I may be justified. The existence of large discrepancies (including phys- The current time has subscript n{tn = nb.t) and we allow both the surface and the
ically impossible K values and corresponding negative exponents), however, suggests average width to vary with time. The finite difference form of Eq. 8.17 is
that the common language has not been found yet. 0

8.4.2 The Combination of the Power Law Assumption


with Interpolation

Returning to the formal material balance (Eq. 8.9) rewritten for the constant injection where we have introduced the constant ~, which is between zero and one. If ~= 1,
case: we take the leakoff rate at the end of the time interval, i.e. we underestimate it. If
~= 0.5, we take the leakoff rate at the middle of the time interval, which is a better
(8.34) choice. The optimal choice is, however, ~ = 0.25 since then the actualleakoff rate is
exactly the average leakoff rate for the first time interval. Our numerical procedure,
one can use different approximations for the last term. Assuming (1) that the value therefore, will be based on
of K is given by
n-l Aj-Aj-1
~ if7]-l ib.t+ (2Sp +4CL~)An-l - 2CL~ 2: .
K= ;' (8.35) j=l ./n - j + 0.25
{
3' if7]-O An =--------------------------~~~-------------
+ 2S + 4Ccv'fV
Wn p
(8.38)
-- ._ ...._ .. -- --_ .._-- __
.. .... ---------.---.~

180 Material balance


Differential material balance 181

where Ao = 0, and we have to know the law according to which the width grows. Table 8.6 Opening-time distribution factor K and
Of particular interest is a simple power-law-type width growth with exponent , estimated power law exponent of

Wj = We C~tr = We (:er (8.39)


fracture surface growth, a. for the data
of Table 8.5
A,
in 1000 m2
K, 01,

because with = 0 it contains the constant-width problem as a special case, the


solution of which is known exactly. In the following we consider a problem with 0 44_3 1.37 0.92
input data similar to the specification of the comparison study [1] to arrive at useful 0.25 43.7 1.48 0.67
conclusions. 0.5 43.1 1.59 0.47
0.75 42.5 1.72 0.28
41.8 L87 0.12
Example 8.5 Material Balance with Power Law Width Growth
=
The specific design problem of Example 8.1 corresponds to i 0.0662 m3/s (25 bpm and the corresponding estimate of the power law exponent of the fracture surface growth
per wing), t, = 12000 s (200 min) and CL =
9.84 x 10-6 rn/s1/2 (0.00025 ft/minl!2). from the numerical solution of the equation
Assume that the final width is w. = 0.015 m. Calculate the fracture surface assuming
different width-growth exponents, 4J = 0, 0.25, 0.5, 0.75, and L aJ]rnOl) _ Kc = D.
n~+a)
Solution
Table 8.6 shows the results. As seen from the table, the constant width ( = 0)
and the moderate width growth (4J = 0.25) result in K factors lying indeed between
Applying the numerical procedure of Eq. 8.37, we have to decide the number of time
the Nolte [3] bounds. (Furthermore, these results imply that a~y a in Table ,8.4 larger
intervals, n : We apply the simple technique, doubling the number of steps until the
than 0.9 contradicts common sense, if the modelers used indeed Carter s leakoff
result does not change up to three digits. The results are shown in Table 8.5. It happens
equation I.) 0
that 210 time steps are enough in all cases. We can check our result for the = 0 case
since Eq. 8.18 provides the solution to this particular case. The identity of the results
gives further confidence concerning the numerical results.
The important message of the example is that the "constant width during propa- 8.6 Differential Material Balance
gation" assumption provides an upper bound of the created surface if the final width
is known. (Our results do not imply that the surface should be below 44.3 m2 in the The global material balance Eq. 8.1 is a simple bookkeeping for the. whole time
base case of the comparative study because the final width from different models may period and for the whole fracture. Equation 8.17 is already more. det~lled bec~use
be different from We = 0.015 m. However, our results do imply that the quotients in it can account for any time instant. If we think about every location m every ume
Table 8.3 should be below unity, if the models really applied Carter's equation I.) instant, we can formulate an even more detailed material balance. .
For completeness, we calculate the opening-time distribution For generality, we assume that not only the flow rate, .q, and th~ cross sectional
area, Ac, vary both with time and location, but also the height, h (FIgure 8.5). Then,
an elementary volume (control volume) between location x and x.+ ~x changes
during the time interval between t and t + ~t according to the equation

Table 8.5 Created fracture surface, Ac, in 1000 nr' for different exponents volume flowing in-volume flowing out-volume leaking offeevolume increase
of width growth, . Numerical solution (Eq. 8.38) with different
fl. and the exact solution (Eq. 8.18)
q(x)6.t - q(x + ~x)~t - VL = [Ac(t + ~t) - Ac(t)]~x, (8.40)

Numericalwith n, steps where


Exact
n, = 128 n, = 256 n, = 512 n, = 1024 hf CL
0
0.25
0.5
44.1
43.4
42.7
43.6
42.9
42.3
44.2
43.7
43.1
44.3
43.7
43.1
44.3 VL = 2.6.x~t
la
--
..;t=r
dh' + 2~x[hJ(t + 6.t) - hJ(t)]Sp, (8.41)

0.75 42.1 41.7 42.4 42.5 and the surface opening-time, r, is a function of the lateral location, x, the vertical =
41.6 44.1 41.8 41.8 location, h'. (In writing Eqs, 8.40 and 8.41 we emphasize only that independent
variable which is important in the given term.)
182 Material balance
Leakoff as flow in the porous medium 183

The material balance, even in its differential form, is not a closed mathematical
model. It is rather a framework which has to be filled with content using physical
principles beyond the conservation of mass. This will be done in the next chapters.
In particular, Eq. 8.43 with boundary conditions is the basis of the Nordgren-Kemp
model and virtually all pseudo-3D models.
h(x)

Ac(X) 8.7 Leakoff as Flow in the Porous Medium

(;l ----x+L1x
The great advantage of the Carter I fluid loss model is its exceptional simplicity,
especially if the spurt-loss coefficient is considered to be zero. The price of simplicity
is that the form of the leakoff volume vs. time curve is restricted. Flow in a porous
q(x)
medium might be more complicated, and in certain cases a more detailed fluid-loss
model might be necessary.
-4----x
An alternative means to describe fracturing fluid leakoff is to decompose and
model rigorously the controlling phenomena. First, the total pressure gradient from
Figure 8.S Local material balance variables
inside a created fracture to the reservoir, /!'P, at any time during the injection, can
be written as
Nordgren [8J considered an important special case where the height is constant.
Then (8.47)
------..:q...:...CA_:_) + _2h_f_C_L
,:_q...:...CX_+_Lll_:_) + AcCt + /!'t) - AcCt) = 0, C8.42)
f!,x ...;t=r /!'t where /!, Pface is the pressure drop across the fracture face dominated by the filter-
cake, /!,Ppiz is the pressure drop across a polymer invaded zone and /!,Pres is ~he
where the opening-time function is simply the inverse of the length-growth function
pressure drop in the reservoir. This concept is shown in Figure 8.6. In a senes
x f(t).After taking the limits /!'t --'>- 0 and Lll --'>- 0 Eq. 8.42 becomes
of experimental works (Mayerhofer et at. [101; Zeilinger, et at. [11]) using typical
aq + 2hfCL + aAc = o.
hydraulic fracturing fluids (e.g. borate and zirconate crosslinked fluids) and cores
(8.43) of permeability less than 5 x-iS m2 (5 md), no appreciable polymer invaded zone
ax ~ at
was detected. So, the second term in the right-hand side of Eq. 8.47 can be ignored.
Considering the material balance, there are two boundary conditions. (Additional
boundary conditions may arise if other variables are introduced.)
(1) At the wellbore, the flow rate in the fracture is the injected flow rate

at x = 0: qtx, t) = iCt), (8.44)


which is the material balance for the wellbore connection.
(2) The material balance at the tip is

at x = x f : q(x, t) = U f (t)Ac (x, t) + 2uf(t)hf(x, t)S p' (8.45)

where the tip propagation velocity, uf, is defined as the growth-rate of the fracture
length
dxf
uf(t) =-. (8.46)
dt '/ '{ Invaded
zone
Equation 8.45 is the so-called Stefan's boundary condition, somewhat overlooked by Filtercake
Nordgren [8] but clarified by Kemp [9]. It shows that the flow rate at the tip is used
Figure 8.6 Fracture filtercake, invaded zone, reservoir and respective pressure gradients for fluid
by creating new volume and by spurt loss occurring while creating new surface. leakoff
-------- ~.... ---- .... ------------------------------

184 Material balance


Leakoff as flow in the porous medium 185
I:I0wever, this is not th~ case when using linear gels in higher-permeability forma-
where K2 is the rate of pressure increase (given in Pals) and A is the retardation time
nons, where ~ polymer Invaded zone is likely to be formed. This would be observed
(given in seconds). Except for very early time, the term itt, is much larger than the
for exa~p~e 10 fracspack treatments. In this section, it will be assumed that filter-
second term and therefore Eq. 8.50 becomes simply (Mayerhofer et al. [13])
cake building fracturing fluids are used in relatively lower permeability formations,
and t~us, only the filter-cake and the reservoir terms in Eq. 8.47 will be considered.
Equation 8.47 reduces to RD =
vr;fi. (8.51)

D.p(t) = !).Pface(t) + !).Pres{t). (8.48) The value of Ro (expressed in m ") can be obtained either in the laboratory (where
it will be a weak function of the selected characteristic time, te) or, preferably, in the
8.7.1 Filter-cake Pressure Drop field from an injection test, where we can select the characteristic time to be equal
to pumping time, t., as will be described in Chapter 9.
The filter-cake pressure term has been given by Mayerhofer et al. [12] as Figure 8.7 shows schematically the evolution of fracture area at different times
while fresh filter-cake is deposited on the older walls. At the fracture tip, the actual
lr/I.Ro deposition of the filter-cake is delayed which may cause considerable fluid loss. The
!). Pface= --;;:-RDq, (8.49) transient leakoff model implies an average filter-cake resistance for each evolving
fracture area. In fact, we can rewrite Eqs. 8.49 and 8.51 into the form
wh~re /.l f is the filtrate viscosity, Ro is the dimensioned, final and characteristic
reslstanc~ of th~ filt:r-cake, :,hich is reached during a characteristic time t.. Using
the _Kelv1O-VOIgt :l.scoelastlc model for the description of the flow through a q= ( D. Pface..;t;) A-,
1 (8.52)
lr/.lfRO ..ji
continuously depositing fracture filter-cake as depicted in Figure 8.7, Mayerhofer
et al. [12J have shown that the dimensionless filter-cake resistance is where the quantity in parentheses can be interpreted as a "pseudo" leakoff coefficient.
In this approach, the time is measured from the start of the leakoff (or opening of
the surface). If the surface is opened during a longer period, the time is measured
(8.50) from the opening of the first element of the surface.

8.7.2 Pressure Drop in the Reservoir


The pressure drop in the reservoir can be tracked readily by employing a pressure
transient model for injection into a porous medium from an infinite conductivity
fracture. For this purpose, known solutions are available in the petroleum engineering
literature. The only additional problem we face is that the surface area is increasing
during fracture propagation. Therefore, for every time instant we have a different
fracture length which, in turn, affects the computation of the dimensionless time.
The general solution for the injection of fluid through a fracture of any specified
fracture conductivity into a fissured (w =1= 1) or homogeneous (w = 1) reservoir has
been presented by Cinco-Ley and Meng [14]. Assuming segments of equal length
withxDj located at the jth segment, the equation can be written (in Laplace space) as

Pres.D(s) - ! t
i-I
ZhDi(S) LtD''''! [KO(XDi-x')ySj(S5 + KO(XDj + x'h/sf(s)]
:AD,
dx'

Figure 8.7 Propagating fracture with growing filtercake (8.53)


------_._------------------------- --------.----------------------------~----~---

186 Material balance


References 187

Wr~ting this .equation for every fracture segment, a system of n equations is obtained The inverted form of Eq. 8.59 is in wide use in petroleum engineering and is
which con tams n + 1unknowns. The first n unknowns are the dimensionless flow simply
=
rates per unit- of fracture length, qfDj(s), i 1, ... , n, and the (n + 1)-th one is Pres.D = ..j7iiD;f. (8.60)
Pres,D(S).One additional equation results if we recall that the flow leaving the fracture
has to flow through the filter-cake. Since this leakoff rate is used to create the In general Pres.D can be obtained from solving the system of Eqs. 8.53 and 8.54 or
dimensionless variable for the given time instant, we have using a type curve [14].
According to Duhamel's principle, the linearity of the diffusivity equation allows
n .6..x
L qfDi(S) = -.
i=l S
(8.54)
the application of the superposition theorem as a sequence of constant rates where
the rate history is known and the only unknown is the actual leakoff rate from one
wing, qn:
For transient interporosity flow,
(8.61)

f(s) = (J) + (1- W)AfD~ tanh (~), (8.55)

while for pseudosteady-state flow it is 8.7.3 Leakoff Rate from Combining the Resistances
(Ehlig-Economides et al, [6J)
f(s)=w+ (l-w)Af
(8.56)
(1- w)s + Af Substituting Eqs. 8.61, 8.49 and 8.51 into Eq. 8.48 we obtain

The Cinco.-Ley and Meng [14J solution follows the great tradition of double porosity
systems pioneered by Warren and Root [15] using (l) as the dimensionless fissure t:l.p(tn) ~ fRo
= JT:J.L fin
-qn +~ IL ~
L./qj - qj-l)PD(tn - tj-l), (8.62)
storativity: n te JT: fb f j=l

(J) = etf 4>fctf and a simple rearrangement yields


(8.57)
erf + 4>macima = (el) /
and A as the interporosity flow coefficient

(8.58)
(8.63)
In Eqs. 8.53 to 8.58"
Equation 8.63 can be used in a hydraulic fracture model. It allows determination
kfbf. th dim . of the leakoff rate at time instant t; if the total pressure difference between the
(kf bf )D = -k-- IS e ensionless fracture conductivity fracture and the reservoir is known, as well as the history of the leakoff process. The
f~f '
dimensionless pressure solution, PD(tn - tj-1), has to be determined with respect to
AfbhmaVb a dimensionless time which takes into account the actual fracture length at tn. The
AfD = V ma = Afmahma is the dimensionless fracture network area,
above leakoff model will be used in the next chapter.
kma
J.L(el)ma x} .
TfmaD =
___
kfb hrna
2 IS the dimensionless matrix hydraulic diffusivity, References
j.J.(4)ci)r
1. Warpinski N.R., Mosochovidis, Z.A., Parker C.O. and Abon-Sayed, I.S.: Comparison
4>f is the fissure porosity, and clf
is the fissure total compressibility. Study of Hydraulic Fracturing Models: Test Case-GRI-Staged Field Experiment No. 3,
For w =1 (homogeneous reservoir) and for kfb f ~ 00, the system of Eqs. 8.53 SPE Production & Facilities, 9(1), 7-16,1994.
and 8.54 has a well known limiting solution 2. Howard G.C. and Fast, c.R.: Optimum Fluid Characteristics for Fracture Extension,
Drilling and Production Prac., API, 261-270, 1957 (Appendix by E.D. Carter),
_ JT: 3. Nolte, K.G.: Determination of Proppant and Fluid Schedules from Fracturing Pressure
Pres,D = 2s3(2 . (8.59) Decline, SPEPE, (July), 225-265, 1986 (originally paper SPE 8341, 1979).
1B8 Material balance

4. Nolte, K.G. Fracturing-Pressure Analysis. In Recent Advances in Hydraulic Fracturing


Gildly, J.L et al. (ed.) Monograph Series, SPE, Richardson, TX (1989) Vol. 12 Chap. 14.
5. Abramowitz, M. and Stegun, I.A. (ed.).: Handbook of Mathematical Functions (9th ed.),

7.
Dover, NY, 1989.
6. Valko, P. and Economides, MJ.: Fracture Height Containment With Continuum Damage
Mechanics, paper SPE 26598, 1993.
Meng, Hai-Zui: The Optimization of Propped Fracture Treatments, in: Economides, M.J.
9
and Nolte, K.G.: Reservoir Stimulation (2nd ed.), Prentice Hall, Englewood Cliffs, N.J,
1989. COUPLING OF ELASTICITY,
8. Nordgren, R.P.: Propagation of a Vertical Hydraulic Fracture, SPEJ, (Aug.), 306-314,

9.
1972; Trans. AIME, 253, 1972.
Kemp, L.F.: Study of Nordgren's Equation of Hydraulic Fracturing, SPE Production
FLOW AND MATERIAL
10.
Engineering, (Aug.), 311-314, 1990.
Mayerhofer, M.1., Economides, M.1. and Nolte, K.G.: Experimental Study of Fracturing
BALANCE
Fluid Loss, paper CIM/AOSTRA 91-92 presented at the Annual Technical Conference
of the Petroleum Society of CIM and AOSTRA, Banff, April 21-24.
11. Zeilinger, S., Mayerfoher, M.1. and Economides, M.J.: A Comparison of the Fluid-Loss
Properties of Borate-, Zirconate-Crosslinked and Non-Crosslinked Fracturing Fluids,
paper SPE 23435, 1991.
12. Mayerhofer, MJ., Economides, M.J. and Nolte, K.G.: An Experimental and Funda-
Engineering models for the propagation of a hydraulically induce~ .fracture comb~ne
mental Interpretation of Fracturing Filtercake Fluid Loss, paper SPE 22873, 1991.
13. Mayerhofer, M.1., Economides, M.1. and Ehlig-Economides, C.A: Pressure Transient elasticity, fluid flow, material balance and (in some cases) an additional propaga.tlOn
Analysis of Fracture Calibration Tests, Paper SPE 26527 presented at 68th Annual criterion. In this chapter we derive the basic models and examine the assumptions
Technical Conference and Exhibition, Houston, Texas, 3-6 October 1993. and simplifications of the individual approaches. . . .
14. Cinco-Ley, H. and Meng, H.Z.: Pressure Transient Analysis of Wells with Finite Given a fluid injection history, a model should predict the evolution With time
Conductivity Vertical Fracture in Double Porosity Reservoirs, Paper SPE 18172 of the fracture dimensions and the wellbore pressure. For design purposes, a rough
presented at the 63rd Annual Technical Conference and Exhibition held in Houston, description of the geometry might be sufficient, and hence, simple models, predicting
TX, October 2-5, 1988.
length and average width at the end of pumping, are very useful. Indee~, the le~gth
15. Warren, J.E. and Root, P.1.: The Behavior of Naturally Fractured Reservoirs, SPEJ,
(Sept.), 245-55, 1963; Trans. AIME, 228, 1963. is an important variable from the production point of view,. and so IS th~ Width
16. Ehlig-Economides, CA., Fan, Y. and Econornides, M.J.: Interpretation Model for Frac- which provides the potential to establish conductivity b~ placmg. proppant into the
ture Calibration Tests in Naturally Fractured Reservoirs, Paper SPE 28690 presented fracture. Models which predict these two dimensions while the third one - fracture
at the International Petroleum Conference & Exhibition, Veracruz, Mexico, October height - is fixed are referred to as 2D models. Even when the fracture. height .is not
10-13,1994. fixed a priori, but is postulated to have a circular surface, the model IS considered
to be 2D.

9.1 Width Equations of the Early 20 Models


9.1.1 Perkins-Kern Width Equation

The elegantly simple approach of Perkins and Kern [1] envisions the ~acture ~s
shown in Figure 4.2. The model assumes that the condition of plane. strain ~olds m
every vertical plane normal to the direction of propagation, but - unhk~ the ngorous
plane strain case - the stress and strain state is not exactly the sam~ m subsequ~nt
planes. In other words, the model applies a quasi-~Iane ~train a:'sumptJon. Neglecting
the variation of pressure along the vertical coordinate, It considers the net pressure,
.t. as a function of the lateral coordinate x: As ,:"e saw .in Chap~er 2, the constant
pressure gives rise to an elliptical cross section With maximum WIdth
190 Coupling of elasticity, flow and material balance Width equations of the early 2D models 191

2hfPn of Eq. 9.3 implies


wo=-p

where hf is the constant fracture height and E' is the plane strain modulus. The
(9.1)
~=Jl-
Ww,o
x,
xI
(9.S)

maximum width, wo, is a function of the lateral coordinate. At the wellbore it is


denoted by Ww,o. the lateral component of the shape factor is
From Chapter 5, the pressure drop of a Newtonian fluid in a limiting elliptical
cross section is given by (9.9)
t::.p 64M
(9.2)
L 7IW~h/ = and hence
14 7r
where J1-is the viscosity of the fluid and q is the flow rate at the lateral coordinate x. Y = 7r45 = '5 = 0.628. (9,H)
To simplify the treatment, Perkins and Kern [1] approximated the flow rate with the
injection rate (q ~ i) where i (as before) is for one wing. In reality, q < i, not only Combining Eqs, 9,6, 9,7 and 9.10, the average width is given by
because part of the fluid leaks off but also because the increase of width "consumes"
another part of the fluid. In fact, if there is anything that is more or less constant
along the fracture, it is not the flow rate but rather the flow velocity. Nevertheless, if (9.11)
the Simplification is accepted and Eq. 9.1 is substituted into Eq. 9.2, the following
differential equation is obtained for the pressure:
It is enlightening to couple the Perkins-Kern width equation with a simple material
balance valid for the constant-injection-rate/no-leakoff case. Then,
dp SJ1-iE'3
(9.3)
dx = -7rhj~'
. _ 5127r3) 1/4 ( J1-lx
. 5fhf4) 1/4 ( . 5 h4I )
J1-IXf 1/4
it = wXfhl =
(
--
625
---
E'
= 2.24 ---
E'
, (9.12)
To obtain a solution, one has to specify the net pressure at any location. Perkins
and Kern [11 postulated that the net pressure is zero at the tip. Integrating Eq. 9.3
between the wellbore and the tip yields from which the length growth is obtained as

4
Pn,w - 0 =
4 32E'3 uixI
h4
7r I
(9.4) x = ( 625 )
f 5127r3
1/5 (iJ1-hj
3 E' ) 1/5 t4/5 = 0.524 (iJ1-h}
3 E' ) 1/5 t4/5, (9.13)

and hence, by combining Eqs. 9.1 and 9.4, The maximum fracture width at the wellbore can be expressed by combining Eqs, 9.6
4 512J-Lixf and 9.13 as
w (9.5)
w.o -- --'--~
7rE'
= (2560)
liS (2_!...!::_ ) 1/5 tl/5 = 3.04
('2_!__!!:_ ) 1/5t 1/5, (9.14)
Taking the fourth root, maximum wellbore width is given by ~o ~ E~ p~

_ (512)1/4 (J1-iXf) 1/4 (J-LiXf) 1/4 (9.6) and by combining Eqs. 9.14 and 9.1, the net pressure is given by
W 0 - - -- :::3.57--
w, n E' E'
Equation 9.6 is the Perkins-Kern width equation. Knowing the maximum width at
Pn,w
= (80)1/4 (E'4J1-P)
2 h6
115 t1/5 = 1.52 (E'4J-Li2)
h6
1/5 t1/5. (9.15)
the wellbore, we can calculate the average width, multiplying it by a constant shape 7r f I
factor, y,
w= YWw,o, (9.7) Thus, in the no-leakoff Perkins-Kern model, fracture length grows with the ~
power of time and the maximum width (and average width) grows with the! power
The shape factor contains tt/4 because the vertical shape is an ellipse. Also it contains of time. More importantly, the pressure is an increasing function of time, growing
another factor which accounts for the lateral variation of the width. Since the solution with exponent !.
-_ .... _---_ .... _-------------

Width equations of the early 2D models 193


192 Coupling of elasticity, flow and material balance

From Eqs. 9.18 and 9.21,


9.1.2 Geertsma-deKlerk Width Equation
336fJ,ix} .
Ww = , (9.22)
The first model of hydraulic fracturing, elaborated by Khristianovich and Zheltov 7rE'hf
[2,3], was based on another geometric picture, shown in Figure 4.3. Those authors
assumed that the fracture opens with the same width at any vertical coordinate within therefore the wellbore width is given by
the fixed height, hf. The underlying physical hypothesis is that the fracture faces
slide freely at the top and bottom of the layer. The resulting fracture cross section . _ (336) 1/4 fJ,IXj
") 1/4 = ( fJ,IXj
''') 1/4
(9.23)
is a rectangle. The width is a function of the coordinate .r. It is determined from the Ww - ( 3.22 E'h
tt Phf f
plane strain assumption, now applied in (every) horizontal plane. The Khristianovich
and Zheltov [2,3] model contained another interesting assumption: the existence of It is convenient to refer to Eq. 9.23 as the GDK equation. If, however, the concept
a non-wetted zone near the tip. of horizontal plane strain approximation is to be emphasized, it is usual to speak
Geertsma and de Klerk [4] accepted the main assumptions of Khristianovich and about the KGD model or KGD view of the fracture. In this case the shape factor
Zheltov [2,3] and - using some innovative mathematical techniques - reduced it has no vertical component Its horizontal component is simply it /4 because of the
into an explicit formula. First, we provide a simplified derivation. elliptical shape of the constant-pressure fracture. Thus, the shape factor is given by
Recalling the plane strain solution of Chapter 2, a constant net pressure, Pn' would
tt
cause a fracture width at the wellbore given by Y =-
4
= 0.785, (9.24)

r-r ::
4x/Pn (9.16) and the average width is simply

(Note that for the KGD model the width at Ww = w,,.o.) Assume again that the _ , ) 1/4 ( fJ,IXf
211r~ . 2) 1/4 ( IJ.IXf
. 2) 1/4
(9.25)
flow rate is everywhere equal to the injection rate, i, Since now the flow channel is W = ( 16 E'hf = 2.53 'hf
rectangular, the total pressure drop from the wellbore to the fracture tip (assuming a
Newtonian fluid) can be calculated from (see Table 5.2) If we compare Eq.9.25 with Eq. 9.11, we can calculate the ratio of the

Pn.w- Pn.tip
12fJ,ixf
= -h-- ( 1
~
loX! 3"1 dx ) . (9.17)
Geertsma-de Klerk width to the Perkins-Kern width:
f xf 0 w
WGK = (21 X 625)1/4 (2xf) 1/4 = 0.95 (2xf)I/4; (9.26)
Now suppose that WPK 32 x 512 hf hf
(1) the pressure equals the wellbore pressure almost everywhere and hence, in
Eq, 9.16, the average pressure can be substituted by the wellbore pressure: therefore the two equations give nearly the same average width for a fracture with
equal vertical and horizontal dimension. In Chapter 4, it was establish.ed that for
4Xf Pn.w
(9.18) 2x f < h f the horizontal plane strain assumption (KGD) is m?re appropnate and for
WW= E' , 2xf > hf the vertical plane strain assumption (PKN) is phys~caUy m?re a~ceptabl~.
Equation 9.26 shows that the "transition" between the models IS essentially smooth.
(2) the net pressure at the fracture tip is zero
As we did for the Perkins-Kern width equation, now we couple the Geertsma-de
Pn.tip = 0, (9.19) Klerk width equation with a simple material balance valid for the constant-injection-
rate/no-leakoff case. Then,
(3) the average value of 1/w3 can be obtained from its value at the wellbore multiplied
by a constant which is postulated to be 7j7r (with explanation later); thus, . _ (211r3 1/4 ( fJ,lxfhf
. 6 3) 1/4 _ . 6 h3f ) 1/4
( fJ,IXf
(9.27)
It = wxfh f = 16) ----p- - 2.43 '

C~loX! :3 ctx) = ~~.' (9.20)


from which the length is calculated as
Combining Eqs. 9.17 to 9.20, we obtain
1/6 ( 3 ') 1/6 ( '3 E' ) 1/6
= (~) iE t2/3 ::::0.539 _, _ t2/3. (9.28)
(9.21) xf 211r3 JLh} J.Lh}
----,---,- -- ----- ------------- "-------_
---~------~--------------

194 Coupling of elasticity, flow and material balance


Width equations of the early 20 models 195

The wellbore width can be calculated now from Eq. 9.22 and 9.28 as
which is a good approximation of Eq. 9.31 for Xo -4- xf. Also they stated that
Eq. 9.33 can be approximated by
i3u.
W
w
= ( 5376)
--7(3 1/6 ( i3 j.L ) 1/6
--
E'h3
tI/3 = 2.36 (
__
E'h3
) 1/6
ti/3. (9.29)

R
! f
_ 2_xo
Substituting Eqs. 9.28 and 9.29 into Eq. 9.21 the net pressure is obtained as P- amin ::::;-P- . (9.35)
n xf

Pn,w = (~~) 1/3 (E'2j.L)1/3rI/3 = 1.09(E'2j.L)1/3t-i/3. (9.30) From Eqs. 9.34 and 9.35, they obtained our Eq. 9.16. Also, our Eq. 9.20 was not
used by those authors, who used instead
Thus, in the no-leakoff Geertsma-de Klerk model the length grows with the ~
power of time; the wellbore width (and average width) grows with the! power of
time. The net well bore pressure behavior deserves further attention. (9.36)
Equation 9.30 shows that the net pressure decreases with time. This is a well
known result of the model. In massive hydraulic fracturing, however, the net pressure
is more often increasing with time. Even more startling is the (less well known) other
consequence of Eq. 9.30: The net pressure does not depend on the injection rate. This which was derived by the "visual examination of a plot of the left-hand-side" of
contradicts the daily experience of fracturing workers. Eq. 9.36. The plot was created by substituting Eq, 9.31 into the left-hand side and
Some notes regarding our Eq. 9.23 are necessary here. We gave a possible "deriva- calculating the integral numerically.
tion" emphasizing the "symmetry" with respect to the Perkins-Kern model. The three Our factor of 7/rr in Eq. 9.20 can be "explained" if we assume that the length of
assumptions leading to Eqs. 9.18 to 9.20 were not stated by Geertsma and deKIerk. the unwetted zone at the tip is always 0.0877 x], because then (using the elliptical
On the contrary, they postulated that the tip net pressure equals minus one times shape)
the minimum far-field stress, i.e. there is vacuum near the tip. They considered a
"zipper" crack with piecewise constant pressure having the value p in the interval (9.37)
from x = 0 to a certain value x = Xo, and the value zero from x = Xo to the tip,
x = x!. The width as a function of the location for that case was derived in our We do not believe that this assumption has strong basis. We gave our version
Example 2.6 and takes the form
of the "derivation" only because we wished to show that the Geertsma-de KIerk
width equation is not inherently connected with the "zero absolute pressure at the
(9.31) tip" assumption. Of course the only authentic derivat.ion of the Geerts~a-de KIerk
equation is the one given by those authors, and the interested reader IS referred to
where their original paper [4].
q1 = (x} - x~)1/2

q2 = (x} - x2)1/2 9.1.3 Radial Width Equation


(9.32)
A closer look at Eq. 9.26 may convince us that ahorizontal plane strain (KGD-typ~)
_ { (x6 - x2)1/2, if x :5 Xo
-xo2)1/2 ' ifx>xo
q3 - (2 width equation could be postulated with the requirement that the constant factor m
X Eq. 9.26 be unity. Such a width equation would have the form
and
Xo = x! Sin
. npn.w . _'--
= x! sm n(p - O"min)
_ 224
. 2) 1/4
= 2.66 ( . 2) 1/4
2(Pn,w - Pn.tip) 2p
(9.33) - - 21/4 X
w- . ( j.LIXI
E'h!
j.LIX!
E'lt]
(9.38)

Starting from Eqs. 9.31 and 9.32, Geertsma and deKIerk made a series of approxi-
mations. They described the width at the wellbore by A favorable consequence of the above is that it can be generalized for a radi~lly
propagating fracture, i.e. xI = h! /2 = R. In this case the same average ~acture WIdth
(9.34) will be calculated from (1) a PK width equation with given xI, (2) a honzontal plane
strain width equation with h! = Zx] and (3) a radial width equation with R = X!.
196 Coupling of elasticity, flow and material balance Algebraic (20) models as used in design 197

The corresponding radial width equation is where the constant factor, 3.27, is derived from an interesting limiting result of
Nordgren [6], who himself has never advocated its use in such context.
_w=2.24 (fLiR)
E! 1/4
(9.39)
(In the petroleum engineering literature, Eq. 9.40 is often written in terms of
(1 - v)/G [= 2/E'] and total injection rate for two wings. Then the factor 3.27 has
to be reduced by 41/4 and it becomes 2.31.)
(Notice that the same average width means smaller volume because the area of the Using the shape factor as given by Eq. 9.10, the average width is obtained from
circle is less than the area of the square.) Depending on the author's preference in
applying analogy, different constants are used in the literature. Geertsma [5] gives
1.32 and 1.62, derived from some additional considerations. (9.41)
The constant-injection!no-Ieakoff case for a radial model results in the following:
The radius, R, varies with the ~ power of the injection time, and hence, the fracture and the constant in Eq. 9.12 becomes 2.05 instead of 2.24.
surface area grows with the ~ power of time. The width grows with the ~ power of For reasons that are not delineated here, several other shape factors are also in
time (since the volume is a linear function of time). The pressure is proportional to use by certain authors, e.g. 3n"/16 (see Economides et al. [7], p. 434) or 7[2/16 (see
the minus ~ power of time, thus giving the same characteristic pressure behavior as Nierode [81, p. 400).
the KGD modeL Here, however, the decreasing pressure seems to be more logical The letter "C" in the name PKN-C denotes that we use the Carter II solution of
because radial fracture growth implies the absence of growth barriers in the vertical the material balance. Expressing the fracture length from Eq. 8.18
direction, i.e, less resistance to growth.
xf = (w+.,2S {J )i [.,exp({3-)erfc({3)+ 2{3]
'- - 1 where {3=
2CL.fiii
, (9.42)
4Cinhj ..;7[ w+2Sp
9.2 Algebraic (20) Models as Used in Design
we obtain a closed system of equations from which either the length (simulation
At this point we have sufficient information to present some widely used design mode) or the time (design mode) can be easily determined using a numerical root-
models. It is assumed, that hf, E', i, u, CL and Sp are known, and two problems finding method. The corresponding net wellbore pressure is determined from
are considered. In design mode, a target length xf is given and the pumping time, t.;
is determined from the model; final net wellbore pressure, wellbore width, average '
PII,\>" = 2hf Ww.O (9.43)
width and fluid efficiency are useful byproducts of the calculations. In simulation
mode, the pumping time is given, and the length is determined from the model; the
same byproducts are generated.
The governing idea behind these algebraic models is to assume the"validity of a Example 9.1 PKN-C Simulationand Design
width equation such as Eq. 9.11 (even if the leakoff cannot be neglected) and combine
Determine the created fracture length, maximum wellbore width, average width and
it with a suitable form of the material balance. Remember that the derivation of these
fluid efficiency for the data used in the comparative study by Warpinski et al. [9} for
simple width equations involves the assumption of uniform flow rate everywhere in a pumping time of 200 min (simulation rnode.) The input data (changed into 5I units)
the fracture. As a fracture changes width and/or loses fluid through the fracture faces, are given in Table 9.1. Notice that the injection rate, i, is defined for one wing. (The
the assumption of constant flow rate along the lateral coordinate is not correct. Once plane strain modulus, E', was obtained from E' = /(1- \}2), where E is the Young's
we accept this theoretical inconvenience (which seems to affect the numerical values modulus and l! is the Poisson ratio as given in Table 2.1).
calculated from the models, but not their characteristic behavior) the treatment is Repeat the calculation in design mode if the target length is 1000 m (3280 ft).
rigorous and the results are very useful.
Table 9.1 Input data for Example 9.1

9.2.1 PKN-C 9.84 x 10-6 m/sl/2 0.00025 ft/minl/2


Om o in
A considerable part of the petroleum engineering literature considers Eq. 9.6 some- 51.8 m 170 It
what inaccurate and uses instead an "improved" constant: 6.13 x 1010 Pa 8.89 x 106 psi
0.2 Pa-s 200 cp
0.0662 m3/s 50/2 = 25 bpm
fLiX
E/ ) 1/4 12000 s 200 min
Ww,O = 3.27 ( (9.40)
198 Coupling of elasticity, flow and material balance Algebraic (2D) models as used in design 199

Solution 9.2.2 KGD-C


Equations 9.41 and 9.42 can be reduced to one equation, eliminating all variables but The procedure for design and simulation with the Geertsma-de Klerk width equation
xf' Introducing the values given in Table 9.1, the equation takes the form combined with the Carter II material balance equation is, of course, similar to the one

2.05 (
0.0662 x 0.2 x x
6.13 x 1010
f)
1/4
0.0662 [2 2{3]
described for the PKN-C model. Traditionally. this model is called KGD and we add
the letter C to indicate the special form of the material balance. Instead of Eqs. 9.40
and 9.41, Eqs. 9.23 and 9.25 are used, respectively. Equation 9.43 is replaced by
Xf = 4{9.84 X 10-6)2 x tt x 51.8 exp(p )erfc({3) + ..fii - 1
'
Pn.w= -Ww (9.44)
where 4xf
2 x 9.84 x 1O-6..jrr x 12000
{3 = 1/4'
2.05 (0.0662 x 0.2 x Xf) Example 9.2 KGD-C Simulation and Design
6.13 X 1010

The solution (to three significant digits) isxf =


1340 m (4400 ft), which can be obtained Repeat the simulation and design calculation of Example 9.1 with the KGD-C model.
from any suitable numerical method. Using Eqs. 9.40, 9.41, 8.5 and 9.43 the necessary
byproducts of the calculation are easily obtained. The results of the calculation are
reported in Table 9.2 in a format used throughout this chapter to ease comparison with Solution
other models.
Equations 9.25 and 9.42 can be reduced to one equation, eliminating all variables but
The calculations in design mode are very similar to the ones in simulation mode.
xr, namely
We substitute the given xf == 1000 minto Eq. 9.41 and calculate the average width
according to ~2 1/4
0.0662 x 0.2 x Xf )
2.53 ( 10 0.0662 2
IV = 2.05 0.0662 x 0.2 x 1000
1/4
= 7.88 X 10-3 m.
x
f
= 6.13 x 10 x 51.8
4(9.84 x 10-6)2 x tt x 51.8 p
+ __p_
[ex ({32)erfc({3)
..fii
_ 1] .
( 6.13 x 1010 )

where
After substitution of the average width, Eq, 9.42 takes the form 2 x 9.84 x 1O-6.jrr x 12000

7.88 X 10-3 x 0.0662 . [ 2{3 ]


f3 = ( 0.0662 x 0.2 x x} ) 1/4 .

1000 = 4(9.84 X 10-6)2 x :rr x 51.8 exp(,82)erfc({3) + ..fii - 1 , 2.53 6.13 x 1010 x 51.8

The results are given in Table 9.4. The design calculation for xf = 1000 m yields
where Table 9.5. 0
{3== 2 x 9.84 X 10-6 v'7fXt .
7.88 X 10-3 The KGD width is larger than the PKN width. Consequently, the KGD length
is smaller (and the desired pumping time larger) than the respective PKN values.
The results are given in Table 9.3. 0 This explains why the KGD model is used from time to time far outside its region

Table 9.2 PKN-C simulation with data in Table 9.1 Table 9.4 KGD-C Simulation
XJ = 1340 m (4400 ft) 'l = 74.1% Xf = 748 m (2450 ft) 'l = 85.7%
ww,o == 1.35 X 10-2 m (0.531 in) P'.IV = 7.98 X 106 Pa (1160 psi) Ww = 2.24 X 10-2 m (0.881 in) P'.W = 4.58 x lOS Pa (67 psi)
IV = 8.48 X 10-3 m (0.334 in) IV = 1.76 X 10-2 m (0.692 in)

Table 9.3 PKN-C design for data of Table 9.1 Table 9..5 KGD-C Design
/, = 8060 s (134 min) 'l =76.5% t, = 1.88 X 10' s (313 min) 'l = 84.7%
Ww.O = 1.25 X 10-2 m (0.494 in) P a ,,., = 7.42 X 106 Pa (1080 psi) Ww = 2.59 X 10-2 m (1.02 in) Pn.w = 5.30 x lOS Pa (77 psi)
w = 7.88 X 10-3 m (0.310 in) IV = 2.03 X 10-2 m (0.800 in)
... -----.. ~---------

200 Coupling of elasticity, flow and material balance


Algebraic (2D) models as used in design 201
of primary validity; engineering intuition and/or other practical considerations may
require a "larger than PKN" width, and the KGD model provides it. The predicted
9.2.4 PKN-a and KGD-a
net pressure for the KGD case is an order of magnitude less than the PKN case, The analytical solution of the material balance assuming power law length growth
indicating that such large KGD fracture is nearly a "perpetuum mobile" needing (Eq. 8.33) can also be combined with any width equation. Instead of Eq. 9.42, we use
almost no energy to propagate.
it

9.2.3 PKN-N and KGD-N (9.46)


Xi = [ a..jirf(a) ] .
iii + 2Sp + 2CL..fi f(3/2 + a)
The Carter II material balance can be readily replaced by the "interpolation between
the Nolte bounds" procedure (Eq. 8.36) shown in Chapter 8. Instead of Eq. 9.42,
We saw in Section 9.1.1 that a = ~for the no leakoff case. It is reasonable to assume
we use
that the exponent remains the same in the presence of leakoff.
(9,45) (Note that when applying Eq. 9.46 for any selected time, t, "neither w: nor the
fracture has to know" how long the injection will last. Equation 9.46 satisfies the
where 1J = hfxfw/it. principle that the future cannot influence the past, and it can be used to calculate the
length at any time during injection without referring to the width or efficiency at the
end of pumping. Unfortunately, the C- and N-type material balances, i.e. Eqs. 9.42
Example 9.3 PKN-N Design and 9.45 do not satisfy this basic requirement.)

Repeat the design calculation of Example 9.1 with the PKN-N model.
Example 9.4 PKN-a Simulation

Solution Repeat the simulation calculation of Example 9.1 with the PKN-a model.

From Eq, 9.41 the average width is given by


Solution
0.0662 x 0.2 x 1000) 1/4 _
W = 2.0 5 ( = 7.88 x 10-' m. Equations 9.41 and 9.46 can be reduced to one equation, eliminating all variables but
6.13 x 1010 XI' Thus,
and hence, Eq. 9.45 takes the form 12000
0.06625l.8
0.0662 x
t -3 XI =
( 1(4

+2 x
1000 x 51.8 = 7.88 x 10 2 OS 0 0662 x 0.2 ~ XI) 9.84 x 1O-6..j12000 x go(~)
. 6.13 X 101

+
6 010-Jr [8 1000 x 51.8 x 7.88 x 10-3 where
.13x1 t:3 0.0662 x I
~ = O.8fi['(0.8) = 1.415.
goes> r(3/2 + 0.8)
+ ( 1- 1000 x 51.8 x 7.88 x 10-3)
J[
]

The results given in Table 9.7 are hardly distinguishable from the ones obtained in
0.0662 x I
Example 9.1: 0
The results are given in Table 9.6. They are almost identical to the ones obtained in
Example 9.1. 0 Note that the results are very similar to the ones obtained with the PKN-C and
PKN-N models. The only reason why we favor the a-type material balance is that it
Table 9.6 PKN-N Design
Table 9.7 PKN-Ct Simulation
t, = 8080 s (135 min) 'I = 76.2%
W = 7.88 X 10-3 m (0.310 in) Xf= 1330 m (4370 ft) lJ = 73.5%
Pn." = 7.42 X 106 Pa (l080 psi)
wlV.o = 1.25 X 10-2 m (0.494 in) o = 1.35 X 10-2 m (0.530 in)
W"'. Pn.w = 7.97 X 106 Pa (1.16 x 1aJ psi)
W = 8.46 X 10-3 m (0.333 in)
-
/

Algebraic (2D) models as used in design 203


202 Coupling of elasticity, flow and material balance

lacks the logical contradiction of incorporating an end-value into the description of Eq. 9.51 takes the form
a dynamic process.
_ [1 +
f-Le - K
(n _ l)n]n
nn
(rr2)n-l
2
(_i
-2h
W f
)n-l
9.2.5 Radial Model

The R-C model can be formulated as follows:


=K [1 + (rr- 1)n]n (25)n-l (_/ )n-l (9.52)
rrn 2 ww,ohf
f-LiR) 1/4 where we have used y = n/5 to relate average fracture width to the maximum
W=2.24 ( - , (9.47)
E' fracture width at the wellbore. .
Substituting Eq. 9.52 into Eq. 9.40 yields the PKN maximum width equation in
R= (W + 2S
4CIIJ
)i [ 2f3 ]
exp(f32) erfc(f3) + -;-- 1 , where f3 = 2CL.J1ii
--- terms of the power law parameters as
W+2S p

(9.48) WW,o
. = 9.151/(2n+2) x 3.98n/(2n+2)
[1 + 214n]n/(2n+2)
n'

'n h1-n ) 1/(2n+2)


and x K1/(2n+2) I f xf (9.53)
(9.49) ( E'
For the KGD model, the results of Section 5.2.2 imply
where the geometry factor relating the average width to the wellbore maximum width
is postulated mostly by analogy. In this book, we accept the value y
Geertsma [5].
The R-N model is the same, except for Eq. 9.48, which is replaced by
given by = Is, f-Le = 2n;1 K c: r r- 2n (~hf 1
(9.54)

and the corresponding KGD width equation is


R2nW
where n = -.-.
It
(9.50)
Ww = 1l.11/(2n+2) X 3.24nl(2n+2) [1: 2nr /(2n+2)
K1/(2n+2)
2)
(.n~fi,
1/(2n+2)

(9.55)
9.2.6 Non-Newtonian BehaVior
There are several ways to incorporate non-Newtonian behavior into the width Example 9.5 PKN-C and KGD-C Simulation Involving Power Law
equations. A convenient procedure is to add one additional equation connecting the Fracturing Fluid
equivalent Newtonian viscosity with the flow rate. Assuming power law behavior,
from Table 5.9 the equivalent Newtonian viscosity is given by Repeat the simulation calculation of Example 9.1 with the PKN-C model and th.en w~tb
the KGD-C model, but instead of constant viscosity, assume a power law behavior WIth

JLe = K [1 + (n -1)n]n (2rruav )n-l ,


g
(9.51)
the parameters of Eq. 5.5 given in Table 9.8.

1m wohf
Solution
where both the average linear velocity, uavg, and maximum width of the ellipse, wo,
are functions of the location. In accordance with the Perkins- Kern model, the' flow Equations 9.53,9.41 and 9.42 can be reduced to the following, eliminating all variables
rate is considered constant and equal to the injection rate. Since the average cross but XI:
sectional area available for flow is Whj. we substitute uavg = i/ (wh f) to obtain the
linear velocity in a representative (average) cross sectional ellipse. The maximum Table 9.S Power law parameters
width of an average elliptical cross section is expressed through the average fracture 0.5
n
=
width using the geometric relation Wo 4W/n. (Note that Wo does not correspond K 2.87 Pa 5.5 0.060 lbf s.5/fI2
to the wellbore but to an "average" location.) After trivial algebraic manipulations,
204 Coupling of elasticity, flow and material balance Differential 2D models 205

Table 9.9 PKN-C Simulation (Power law fracturing fluid, parameters in Table 9.11 PKN-NMB Simulation (Data in Table 9.1)
Table 9.8)
XI= 1320 m (4330 ft) 17 =72.6%
XI = 1200 m (3930 ft) w = 9.86 X
fl., = 00409 Pa- s (409 cp)
10-3 m (0.388 in) = 1.34 X 10-2 m (0.529 in)
W",.O p. 'N =7.95 x 1(1" Pa (1150 psi)
W,,.O =
1.57 X 10-2 m (0.618 in)
17 =76.9% w = 8.44 X 10-3 m (0.332 in)
p; .w = 9.28 x 106 Pa (1350 psi)

is still valid, but now Wn, the average width at time tn, is not known a priori, but
Table 9.10 KGD-C Simulation (Power law fracturing fluid, parame- has to be determined from the PKN width equation (Eq. 9.41). Therefore we have
ters in Table 9.8)
a system of two equations with two unknowns. The system can be reduced to one
Xf = 600 m (1970 ft) w = 2.27 x 10-2 m (0,892 in) nonlinear equation, which can be solved numerically. Omitting the details of the
fl.. = 0.858 Pa . s (858 cp) IJ = 88.6% derivation, the PKN-NMB algorithm is as follows:
W", = 2.88 x 10-2 m (1.14 in) Pn,,, = 7040 X 10l Pa (107 psi) (1) Determine the constant al in the width equation

_W = 2.05 (J,Li)
E'
1/4 1/4
xf = alx f1/4. (9.57)
XI = 4(9.84
wx 0,0662
x 10-6)2 x J'[ x 51.8
[
exp(,81)erfc(fJ)
2fJ
+ ..fii -
J
1 ,
(2) At time t, solve the nonlinear equation
where
fJ = 2 x 9.84 x 1O-6J;r x 12000 alx~/4 + (4CL~ + zs,, =
w '
and

W
;r 1/(Z O.5+2)
= _9.15 x x 3.9S0.5/{2xO.s+2)
[1 + 214
. x.05] 0.5/(2xO,5+21 (9.58)
5 0.5
for x-: (All other variables are known because they correspond to values of previous
X 2.871/(2x05+2) (0 066 2 05 x ::>1.
- 81-05

6.13 x 1010 x 51.8 f


x
) 1/(2xOj-2)
time steps.) Once Xn is known, calculate the new average width, Wn = alx~/4 and
continue with the next time step. (Clearly, a similar algorithm can be derived for the
KGD and the radial width equations.)
The results shown in Table 9.9 indicate a somewhat shorter and wider fracture
(compared to the one of Example 9.1). That is the result of increased viscosity. A new
item appearing in the table is the equivalent Newtonian viscosity, calculated for an
Example 9.6 PKN-NMB Simulation
elliptical cross section which corresponds to average conditions at the end of pumping.
Similar calculations for the KGD-C simulation problem yield the results shown in Repeat the simulation calculation of Example 9.1 with the PKN-NMB model.
Table 9.10. 0

Solution
9.3 Numerical Material Balance (NMB) with Width
Growth Dividing the injection period into 2, 4, 8, 16, ... equal intervals, we check the conver-
gence of the length. The change from 128 to 256 time steps does not cause any changes
By now the reader should not be surprised that several different combinations of the out to three significant digits. Therefore, we accept the results calculated with 256 time
elastic, flow and material balance submodels exist. In this section, we consider an steps. The results given in Table 9.11 are hardly distinguishable from those obtained in
interesting case when the width is given by the PKN width equation (Eq. 9.41), but Example 9.1. 0
instead of Eq. 9.42, the more detailed material balance of Section 8.5 is considered.
Equation 8.37,
9.4 Differential 2D Models
, r:-:"
z6.t = 2C LV t:.t L.J n _.J +~ + 25p(An - An-I)
j=i
Aj -Aj-l
+An Wn - An-I Wn-1, (9.56) The inherent drawback of the algebraic 2D models is that they cannot account for
the variation in flow rate with the lateral coordinate. Nordgren'S [6J material balance
---- ------- - --------_--- ---------- ----_ --------------

206 Coupling of elasticity, flow and material balance Differential2D models 207

was derived in Section 8.6 as The wellbore boundary condition, Eq. 8.44 now becomes
oq 2hfCL oAc
-ax + "It~+-=o
- r at (9.59) at x = 0:
aW6
-=---1,
512/.l.
(9.68)
ax TCE'
and will be used below to account for lateral flow rate description.
where the injection rate may vary with time.
The moving boundary (Stefan's) condition formulated by Kemp [10] is given by
9.4.1 Nordgren Equation
(9.69)
As~u?Iing the quasi-plane strain condition introduced by Perkins and Kern [1], the
elliptic cross sectional area, Ac, of the constant height fracture is given by
where the tip propagation velocity, U [ is the growth rate of the fracture length, i.e.
TChf
AC=4wo, (9.60) U f = dxf I dt. This equation shows that the flow rate at the tip equals the sum of
volume growth rate at the tip plus the rate of spurt due to creating new surface.
where Wo is a function of the lateral location, x. If the cross sectional area varies Another useful form of Eq. 9.69 is obtained by rearrangement:
with time, this is due to the width change and hence,
x =xr rrE'w6 (awo)
aAc = TChf (awo) .
(9.61)
at uf = - 32JLhf(rrwo + 8Sp) & . (9.70)
at 4 at
Whichever form is used, it should be clear that the propagation velocity is an addi-
Similarly, the flow rate at the given location can also be expressed through the
width, firstly rearranging Eq. 9.2 into the form tional variable that is obtained from the system of equations. To obtain a closed
system, another boundary condition is needed which determines the propagation

q
= _T(Wghf
64/.l
(oPn)
ax ' (9.62)
velocity, either directly or indirectly.
Nordgren considered the special case when (1) there is no spurt loss and (2) the
net pressure is zero at x = x f, and hence, the width at the moving fluid front is
and then substituting in the derivative form of Eq. 9.1: zero. If these two assumptions are accepted, the two boundary conditions at the tip
apn
&=2hf
E' (aw&o) . (9.63)
reduce to
Wo =0 (9.71)

The result is and


q = _T(W6E' (awo) _ _ TCE' (awti) (9.64) E'w~ (owo) E' (aW6)
128/.l ax - 4 x 128/.l & ' vr = -32/.lhf & = -96/.lhf & . (9.72)
and hence,
The propagation velocity is finite. Hence, in the Nordgren model, the first derivative
(9.65) of the width (and of the pressure) is infinite (negative). The solution has a singularity
at the tip.
Su?stit:uting ~qs. 9.61 and 9.65 into Eq. 9.59, the following partial differential To understand singularity, the reader may depict a hypothetical width profile of
equation IS obtained for the width: the simple form w = a(xf - x)I/3, where c is a constant. For that profile, the width is
zero at the tip, its derivative with respect to the lateral coordinate is (minus) infinity
TCE'
4 x 128/.l
(aaxwri) _-..;t=r
2
2
2hfC TChf (awo)
+ 4 at
L
' (9.66)
and the derivative of the function w3 = a3(x f - x) is a finite value (in this case
equal to _a3). The solution of the Nordgren equation behaves similarly near the tip
at every time instant: The flow rate is zero at the tip while the propagation velocity
the rearranged form of which is the well known Nordgren [6] equation: is finite.
The system of Eqs. 9.67,9.68,9.71 and 9.72, augmented with the initial conditions
__
128/.l
(o2wti)
ax 2
= 8hf
TC..;t=r
eL +h
f
(awo)
at . (9.67)
at t = 0: Wo = 0 and x f = 0, (9.73)
~--"-'" ----

208 Coupling of elasticity, flow and material balance Differential 2D models 209

and with the implicit definition of the fracture surface opening function, a zero width at the tip. However, together with the basic assumption of the PK view
of the fracture, it means that the net pressure is zero at the tip. We know that the
for every x: x = Xf[T(X)], (9.74)
energy needed to create a fracture is only partly used to increase the elastic energy
constitute the Nordgren and Kemp (NK) model of hydraulic fracturing. The important of the surrounding formation. Another part is dissipated during fracture propagation.
property of the NK model is that it can be written in a dimensionless form where One part of the energy dissipation is considered correctly in ali models: This is the
~h~solution of the dimensionless model is unique. In practical terms, this means that viscous dissipation connected with the flow of the fracturing fluid. The other part
It IS ~no~gh to solve th~ system once (by an appropriate numerical method). If the of the energy dissipated is connected with the irreversible creation of new fracture
solUtl0~ IS repr.esen.ted In the form of graphs or tables or any suitable approximate surface. Within the PK view of the fracture, the zero net pressure assumption also
algebraic equation, 11 can be used for any (constant) injection rate and formation and means that the energy dissipated during the creation of a new surface is totally
fluid data. neglected. In other words, the zero net pressure assumption means that once the
The dimensionless variables are defined as: fluid reaches the tip it can flow further without any resistance; thus it is opening the
rock as "a knife cuts butter".
(9.75) Current evidence of higher fracturing pressures has made it clear that the zero net
where the constants are selected in accordance with Nordgren's suggestion as pressure assumption cannot be valid in general. In most cases, the fracture propaga-
tion is retarded compared to the "knife cuts butter" case. Within the PK view of the
1.5JL ] 1/3 .2 ] 2/3 fracture, and especially in the Nordgren - Kemp model, any accounting for the dissi-
c -rr 2 IJL
1 - [ 128cfE'hj C2 = rr [ 16CEhfE'
pation of energy during the creation of a new fracture surface can be accomplished
by, and only by, dropping the assumption of zero net pressure and using another
c = [ 32i2JL J 1/3
(9.76)
boundary condition, resulting in a positive net pressure at the tip. Chapter 10 is
devoted to a more detailed discussion of this issue. Now we return to the original
3 CtE'hf
model of Nordgren and Kemp, including Eq. 9.80.
The model works well even with variable injection rate. Then, of course, one has to From a strictly mathematical point of view, we need an initial condition. In this
select a nominal injection rate, io, in order to define the constants in Eq. 9.76. The case it is rather trivial, indicating that the fracture is of zero length and width at time
dimensionless partial differential equation is zero.
The resulting system can be solved numerically. Once the solution is known, the
a2w~o 1 awoo fluid efficiency (in fact the dimensionless volume) can be calculated as
(9.77)
axb = ,Jto - 7:D + atD .
One boundary condition assures that the injected fluid rate gets into the' fracture
(9.81)
aw~o
at XD = 0: --
aXD
= --io (9.78)

At the other end, two boundary conditions have to be satisfied. One of them is the In a constant-injection-rate case (i/io = 1), the system has only one solution as was
dimensionless form of Eq. 9.69 with S p = O. It states that the propagation "rate" presented by Nordgren [6] in the form of plots including x fD vs. to and WwO vs. to-
of the fracture (the tip velocity) equals the quotient of the flow rate and the cross (The 11 vs. to curve did not make it into the original paper, causing some difficulties
section, i.e. the linear velocity of the fluid;
in later engineering use of the solution, e.g. see Appendix G in Nierode [8]). Since
there is an uncertainty of how Eq. 9.79 was used (if at all) in the numerical scheme
(9.79) of Nordgren, we do not reproduce the original plots here. Similar plots, however,
will be provided in Chapter 10.
The other boundary condition states that the width at the tip is zero,

at XD = xfO: Woo = O. (9.80) 9.4.2 Differential Horizontal Plane Strain Model


The (sometimes tragi-comic) history of hydraulic fracture modeling can be better A differential 2D model using the horizontal plane strain assumption would require
understood in the light of Eq. 9.80. At first sight it may look rather natural to assume the following ingredients.
210 Coupling of elasticity, flow and material balance Pressure decline analysis 211

The material balance is still Eq. 9.59, but the cross section is defined as leakoff rate, is also unknown a priori, but it is determined by Eq. 8.63. Therefore,
we have a system of equations for every time step.
(9.82) Omitting the details of the derivation, the algorithm is as follows. Suppose we
with time derivative know all the variables up to time tn-I.
<lAc _
at -
hf(aw)at (9.83) (1) First the constants QI and Q2 are determined in the width and pressure equations:

The flow rate expression is derived from the assumption of slot-flow (see Table 5.1) _ = 2.05 (Mi)
W E'
1/4 1/4
xf = QIXf
1/4
' (9.88)

q =_ %h f
12M
(aaxPn) . (9.84) E'3Mi)1/4 1/4 1/4
pwn = 1.63 ( hj xf = a2xf . (9.89)
The coupling between width and pressure is not so simple as in the Nordgren model.
Instead, we have an integral (see Eqs. 2.32 and 2.33): (2) At time t we wish to determine xr. Since we know all the variables up to
time tn-I, the equation
(9.85)

where

(9.86)

and we have to keep both the pressure and the width as variables of the model. As a
result, the system consists of a partial differential equation and an integral equation
(with two integrations in the latter).
The wellbore boundary condition, Eq. 8.44, and the Stefan boundary condition (9.90)
at the tip, Eq. 8.45, remain valid. In this case, the width at the tip will be zero
automatically. In spite of this, the system is not yetdetermined. Again, one additional can be solved for xfn. Note that Eq. 9.90 is highly nonlinear, not only because
tip boundary condition is needed, involving the tip velocity. A strict formulation of the unknown xfn is raised to the power !,
but also because the change from real
the problem in mathematical terms is not available. Limited numerical results and time to dimensionless time involves the actual estimate of x f n- Nevertheless, any
semi-analytical approximations are the subject of current research (Leonach [11]). reliable root-finding method can solve the equation, especially because a good starting
estimate (or lower bound) of xfn is known to be Xfn-l. Once xf" is found, qn is
obtained as either side of Eq. 9.90. Notice that the dimensionless pressure function,
9.5 Models With Detailed Leakoff Description characteristic for the given formation, has to be known. Pressure transient models,
long established in petroleum engineering for both homogeneous and heterogeneous
In Section 8.7 we introduced an alternative means to describe fracturing fluid Ieakoff. formations, are readily available. In using them, the transition from real time to
The models of Sections 9.1 to 9.4 of this chapter can then be reformulated by dimensionless time will involve the actual x f n v valid only for the given time step.
substituting the simplistic Carter I leakoff with a more detailed description of the The parameters of the leakoff model, namely Pr - the average reservoir pressure,
leakoff rate. As an example, consider the derivation of Section 9.2, but substituting f.l - the reservoir fluid viscosity, kfb - the formation bulk permeability, Mf - the
the Carter I assumption with the Ieakoff rate obtained explicitly in Section 8.7. filtrate fluid viscosity and Ro - the characteristic filter-cake resistance (determined
The injection rate is considered constant. We assume that the PKN width equation with respect to the selected reference value of time equal to the injection time, te),
(Eq. 9.41) describes the width at any time instant. The material balance in the n-th have to be known as well.
time interval is given by

(9.87) 9.6 Pressure Decline Analysis


where w", the average width of the fracture in the time interval, is not known a Simulation is one way to apply a given model. It is used when we know the param-
priori, but has to be determined from the PKN width equation, and similarly q", the eters and wish to predict the evolution of fracture dimensions with time. The inverse
--_ _-_ _--------------- -------_.,,,---_.-

212 Coupling of elasticity, flow and material balance Pressure decline analysis 213

problem is calle~ par~meter estimation, when we do not know some of the param- Clearly at ~tD = 0, Eq. 9.91 reduces to the original definition of go(ex). The func-
eters, but try to identify them from known data. Unfortunately, the time evolution tion, g(~tD, ex) should be a monotonically increasing function of the dimensionless
of the fracture dimensions in the formation is almost impossible to observe, even shut-in time (since the fluid leaks off and does not return into the fracture). Following
With the most up-Io-date seismic systems. However, the pressure at the wellbore the reasoning of Section 8.4.1, it can be shown that the g(6.tD, ex) function is the
~an be observed rather accurately. Here we consider a very special model identifica- integral .
non p~oblem, termed pressure decline analysis. It is the valuable engineering means
to denve the parameters of the leak-off process from observation of the pressure 1 (ll+AtD 1 dtD ) dAD.
.
behavior after the pumps are stopped.
g(~tD, ex) =
loo Alia
D
Jto - AI/"
D
(9.93)

~sume that up to the end-of-injection time, te, the injection rate, i, is constant
~nd IS .zero afterwards. The pressure in the wellbore is declining because the fluid Nolte [12,13] gave the closed form of the g(6.tD, ex) function for two distinct values
IS leaking off, and consequently, the fracture faces are approaching each other. A of ex, namely ~ and 1. In Valko and Economides [19] a closed form solution is given
consequence of the narrower width is that the elastic force trying to close back for any ex as
the fracture (i.e. the induced stress) decreases. This effect is seen in the decline of
the :vellbore pressure. Since the whole process is controlled by the leak-off, pressure 4a~ + 2../1 + ~tD x F[~, ex; 1 +ex; (1 + lHD)-lJ
declme a~al~sis has been a primary source of information on the leak-off parameters. g(6.tD, ex) = 1+ 2a (9.94)
If t.helimited descriptive capabilities of the Carter I Ieakoff model (see Eq. 8.8) are
sufficient in a certain application, the pressure decline analysis procedure becomes where F[a, b; c; z] is the hypergeometric function. For definition and properties see
elegant and simple. It culminates in a straight-line fit, a familiar and reliable means Abramowitz and Stegun [20], Chapter 15. The hypergeometric function is a tabulated
of traditional engineering. In the following we give a brief derivation of the method, function, e.g. similar to the sine function, and it is built into up-to-date mathematical
known as Nolte's analysis [12,131, omitting the many minor modifications and software.exactly the same way as the function sine is built into the programming
improvements [14-18] but emphasizing a neglected question: the determination of language FORTRAN. Using Mathematica (Wolfram [21]) one can simply call it by
the spurt loss coefficient together with the Ieakoff coefficient. the name HypergeometricZi-L, For the value of ex = ~(basic PKN case), the g(extD, ex)
function is shown in Figure 9.1.
The volume of the fracture after the end of pumping is given by subtracting the
9.6.1 Nolte's Pressure Decline Analysis (Power Law Assumption) spurt loss volume and the leakoff volume from the injected volume, Vi. Thus,

In Secti?n 8.4 we introduced Nolte's power law hypothesis, which augments the (9.95)
assumptions of Carter (constant leakoff coefficient plus spurt loss) with an a priori
knowledge o~ the f~rm of the solution of the coupled system. According to the power
law hypothesis, dunng the constant-rate injection period the fracture surface evolves 7
with power ex of the time, see Eq, 8.20. The ratio of the leak-off volume to the
product 2CLA~0eis denoted by go(a). The subscript 0 indicates that the ratio is 6
v
tak~n at t = teo Assuming that the fracture surface remains constant after the injection V--
period, Nolte [12,13J extended the definition of the ratio to the whole shut-in period 5 V
,
~~
Ii;
...... ~ L
V~
g(t1tD, ex) = VL(t.+At), (9.91) ~ v
2CLAe0e ~
<l
V
0; 3
wher~ the subscript e refers to the end of pumping and Sr is the time elapsed after V
stopping the pumps (shut-in time). The subscript L in VL denotes "Ieakoff", thus 2
/
V L(t,+At) stands for the volume lost during the injection period plus the volume lost V
during the shut-in period up to the time i, + 6.t, but including only leakoff. (The
total volume lost might be larger if there is a non-zero spurt loss.) The dimensionless o 2 3 4 5 .6 7 8 9 10
shut-in time is simply
t-t e. Dimensionlessshut-in lime.ll.to
6.tD = __
(9.92)
te Figure 9.1 Nolte's g function for ex= ~(PKN Newtonian fluid case)
.. _--- .. _---- ... _---- --- ._---_ ... _-----_ ... ---------.----------

214 Coupling of elasticity, flow and material balance 215


Pressure decline analysis

where Ae is the fracture surface (one wing, one face) at the end of pumping. The and .3 with the height. The volume divided by the area is the average width:
actual fracture volume is, by definition, the product of the constant fracture surface
with the time-varying average width, thus
(9.101)

(9.96)
and hence, we obtain
Combining Eqs. 9.91 and 9.96 results in '
Cf.KGD =-. (9.102)
irXf
(9.97) Finally, for a radial geometry, Eq. 2.69 shows that

Hence, the time variation of the width is determined by the g(tltD, Ct) function, the V 16Rpn
W------ (9.103)
length of the injection period and the leakoff coefficient, but is not affected by the - lR27r - 37r' '
2
fracture area.
The fracture closure process (decrease of average width) cannot be observed thus we obtain
37r'
directly. However, from Chapter 2 we know that the net pressure is directly propor- (9.104)
Cf.Rad = 16R'
tional to the average width, Pn = C fW, simply due to the fact that the formation is
described by linear elasticity theory. (The coefficient C f is measured in Pa/m and Once the proportionality constant, CJ, is related to the geometry of the fracture,
it plays a similar role as the constant in Hook's law. In tbe petroleum engineering we can interpret the observed pressure decline in terms of the fluid-loss parameters.
literature, its inverse, l/cf, is called fracture compliance.) Because of the linearity Indeed, adding the closure pressure, Pc- to both sides ofEq. 9.98 allows us to express
between net pressure and width, we can rewrite Eq. 9.97 in terms of the net pressure: the weJlbore pressure as
cfV; !.)
Pn = T -2cfSp - (2cfCL0e) x g(tltD, ce). (9.98) Pw = Pc + CfV
~ - 2cfSp - (2CfCLyte) x g(tltD,Ct

Equation 9.98 suggests that the pressure variation after the pumps stopped is
=b + m x g(tltD, Ct), (9.105)
strongly connected to the leakoff coefficient. To obtain a useful relationship, we showing that the plot of wellbore pressure vs. the g(tltD, Ct) values results in a
need a closer look at the proportionality constant, Cf. straight line if the fracture area, Ac' the proportionality constant, C I> and the l~akoff
The formulas to calculate Cf are easily derived if we accept that during closure coefficient, CL, do not vary with time. Only when the fracture finally closes Will the
there is no lateral flow in the fracture and the pressure is constant along the fracture pressure behavior depart from the linear trend. . . .
at a given time instant. For the vertical plane strain assumption (PKN-geometry) we Equation 9.105 is the basis of Nolte's pressure decline analysis. The technique
apply Eq. 2.68 to calculate the volume of one wing of the fracture. This is done by requires the plot of the wellbore pressure vs. the values of the g~function, as first
substituting C with half of the height and 8 with the half-length, and then multiplying suggested by Castillo [17]. The g-function values are genera.ted ~lth .the exponent,
the result by two because both wings of the vertical line crack are present The Ct, considered valid for the given model and rheology. A straight line IS fitted to the
average width is the volume divided by the area and hence,
observed points.
For the straight line fit, that portion of the data should be selected which is after the
tt p (hf)2 pumps are stopped but before the fracture faces touch again. To restrict the straight
W::::: ~ = 2 x xf n 2 x _1_ = 1rpnhf, line fit to this period of time, the closure point, i.e. both the time of closure .and the
(9.99)
hfxf ' hfxf 2' corresponding value of the pressure (pc), have to be selected, at least .approxlmately.
In many cases, the closure point is manifested by a clear ~hange m the pr~ssure
and thus, we obtain
behavior, and a plot of the pressure decline curve vs. any SUitable ~ransform~tlo.nof
2' the shut-in time (including the g-function, square root or even straight shut-In time)
Cf.PKN = --. (9.100)
irhf can be used for this purpose. However, a better practice is to obtain the closure
pressure from independent information (see Chapter 3).. .
Similarly, for the horizontal plane strain assumption (KGD-geometry) we apply The slope of the straight line is denoted by m and the intercept by b. ~Irst
Eq. 2.68 to calculate the volume of one wing by substituting C with the half length we concentrate on the slope, which is related to the unknown leak-off coefficient
'---'~"-'---' ---- ---- _._-_ ---_._. __ ._._---_- _---------_.

216 Coupling of elasticity, flow and material balance Pressure decline' analysis 217

according to Eq. 9.105 by Assuming that the g-plot shows a straight line, the Nolte analysis still gives the
CL = (-m)
(9.106)
same answer for the leakoff coefficient even if there is a non-zero spurt loss. The
2.J"i;cf effect of the spurt loss is concentrated in the intercept of this straight line with the
g = 0 axis. From Eq, 9.105, the spurt loss coefficient can be expressed in terms of
Substituting the relevant expression for the proportionality constant (Eq. 9.100), for
the PKN-geometry we obtain the intercept, b, as
S _ ~ _ b- Pc
(9.110)
7Chf p - 2A, Zc]
CL = --(-m). (9.lD7)
4.,fi;E' Thus, for the PKN-geometry the spurt-loss coefficient is given by
Similarly, for the KGD-geometry the result is
(9.111)
CL = 7Cxf (-m) (9.108)
2.,fi;E' '
for the KGD-geometry,
and finally, for the radial geometry the leakoff coefficient is given by
Vi nXf(pc - b)
(9.112)
CL =
8R
(-m). (9.109)
Sp = 2"Cfhf + 2E' ;
37r.,fi;E'
and for the radial geometry,
.Note that the names "PKN geometry" and "KGD geometry" are used, but actually
neither the PKN nor the GDK (KGD) width equations have been utilized in the Vi 8R(pc - b)
(9.113)
derivation. In the expression of the proportionality constant, C f' only basic linear Sp = 2nR2 + 37CE'
elasticity relations are included. In other words, not the PKN model, but only some
Before making use of the intercept, the selection of a closure pressure, Pc- cannot
of its underlying assumptions, are used, and the same is true for the KGD model.
be- avoided. Once the closure pressure is considered known, Eqs. 9.J.ll to 9.113
If anything, from the propagation models themselves, it is only the exponent a that
provide a restrictive relationship between the extent of the created fracture and the
is used (to generate the g-~nction values). The exponent should be ~, ~ and ~
spurt-loss coefficient. At this point two possibilities arise. One possibility is to sacri-
for the PKN, KGD and radial models, respectively. To apply the technique if the
fice the spurt-loss coefficient, i.e. to assume that Sp = O. The other possibility is to
fluid is non-Newtonian, a somewhat different a should be selected. Nolte et al. [16]
derive additional condition( s) from considering the fracture propagation period.
advocate the use of (2n + 2)/(2n + 3) for the PKN geometry, (n + l)/(n + 2) for
the KGD geometry and (4n + 4)/(3n + 6) for radial geometry, where n is the power
law exponent of the rheological equation of the fluid. Fortunately, the estimate of 9.6.2 The No-spurt-lose Assumption (Shlyapobersky method)
the leakoff coefficient is not very sensitive to a small variation in a.
Equation 9.107 deserves further attention. It shows that assuming the vertical Once we restrict the Carter fluid loss model, specifying that the spurt loss coefficient
plane strain assumption (PKN geometry), the estimated leakoff coefficient does not is zero, the intercept of the g-plot provides an estimate of the characteristic length.
depend on unknown quantities since the time of pumping is known, the fracture Indeed, setting the right hand sides of Eqs. 9.111 and 9.112 equal to zero, xf can
height can often be considered equal to the length of the perforations in the well be expressed, and setting the right hand side of Eq. 9.113 to zero, R can be easily
and the plane strain modulus can be assumed with some confidence. To estimate the estimated. Such a procedure (including even the compressibility of the fluid in the
leakoff coefficient, there is no need to know the viscosity of the fluid or the created wellbore) was first suggested by Shlyapobersky et al. [22J.
fracture area. Neither does the exact value of the closure pressure have to be known. Perhaps the most useful result is obtained for the radial case. The no-spurt-loss
The elegance and usefulness of the Nolte technique - at least for the PKN geometry estimate of the radius, Rnsp, of a radial fracture is obtained as
and when only the leakoff coefficient is considered - lies in the independence of
the estirna_tedleakoff coefficient on such uncertain variables as the closure pressure, 3 3E'Vi
and especially, the created fracture length. (9.114)
Rnsp = 16(b - PC)
For the other two models considered, the procedure results in an estimate of the
leakoff coefficient which is strongly dependent on the characteristic dimension (xf Since the radial geometry can be a satisfactory approximation for the majority of
or R). Unfortunately, the characteristic dimension is not known without some model "minifrac" tests, an interesting sequence of parameter estimation can be suggested
calculations involving the still unknown leakoff coefficient. (Shlyapobersky et al. [22]). Once the slope and intercept of the straight line are
---_.- .._---_ ..... _---------

218 Coupling of elasticity, flow and material balance Pressure decline analysis 219

known, first Rnsp is calculated from Eq. 9.114 and then CL from Eq. 9.109, substi- From the g-plot straight-line fit first the instantaneous shut-in pressure, Pw.isi is
tuting the no-spurt-loss estimate of the radius on the right-hand side. The procedure determined as
needs a minimum number of input parameters, namely m, b, Vi, i, E' and Pc. Notice Pw.isi = b + m x go(~) =
b + m x 1.415. (9.115)
that a knowledge of the viscosity of the fluid is not necessary.
A similar technique can be easily derived for the KGD geometry, but fracture Note that pw,iSi is not a physically observed pressure. It is the value ~orres~onding
height should be included in the list of input parameters. The no-spurt-loss assump- to zero shut-in time but read off from the straight line. One can consider this value
tion results in the estimate of the fracture length also for the PKN geometry, but as the optimal estimate of the real instantaneous shut-in pressur~. This value cannot
this value is not used for obtaining the leakoff coefficient because the fracture length be read directly from the wellbore pressure record because the instantaneous pump
does not enter the right-hand side of Eq. 9.107. shut-down creates large fluctuations in the pressure in a very short time period.
Assuming a zero spurt-loss coefficient brings about serious consequences with The average width after the end of injection can be obtained from pw.isi accord-
respect to the propagation model. Once the assumption of no-spurt-loss is accepted, ing to
the final extent of the fracture stems from the pressure decline data alone, and for pW.isi - Pc (b+ 1.415 x m - PC)nhl
the description of the fracture propagation one has to select a model which provides Wist =
Cf.PKN
= 2'
(9.116)
the identified fracture extent (or a somewhat smaller extent if some growth of the
fracture is allowed after the pumps stop). Thus, the Shlyapobersky method leads The average width before the end of injection should be the same, and hence,
almost invariably to the discarding of the fracture propagation models described substituting Eq. 9.116 into Eq. 9.41, we can express the estimated length as
in this chapter, since once the spurt loss is set to zero and the leakoff coefficient
is known from the pressure decline analysis, no additional degree of freedom is
left in these models. The answer to this challenge in the Shlyapobersky method is
XI mb
.
W )4
= ( 2.05 ('.)
tu
= 0.342 ( hi\) (b+ 1.415
J-tV .E
x m _ pc)4, (9.117)

the introduction of a new degree of freedom into the known models of hydraulic
fracturing, utilizing the apparent fracture toughness concept [22]- The concept will where the subscript "mb" stands for material balance estimate. Introducing the length
be discussed in Chapter 10. estimate into Eq. 9.111 provides the optimal estimate of spurt loss for the PKN
model as

9.6.3 Material Balance and Propagation Pressure Estimates of the J-tV2'3) -4 0.785hf(pC - b)
S".mb = 1.46 ( ~5-'- (b + 1.415 x m - pc) + E' . (9.118)
Spurt Loss hfte

It is not necessary to assume that the spurt-loss coefficient is zero, but then the Accepting this estimate of the spurt loss to.gether ,:i~h t~e esti~ated leakoff
estimation procedure should include some independent information. coefficient has the following consequence. Dunng the mjecnon period, the p~
The additional information cannot come from the shut-in period only, because that propagation model will arrive at length Xfrnb- which, combined with the two ~uld
part is described by the straight line on the g-plot and any reasonable combination of loss parameters, will force the pressure decline curve to follow the path determined
S p and X I (or R) satisfying 9.110 will result in the same pressure decline curve. The by the two parameters of the straight line, m and b. If the observed data fo~lows a
independent additional information has to be connected with the injection period. straight line on the g-plot, the combined simulation of the .fracture propagation and
Here we suggest two alternative approaches. The first one determines the spurt pressure decline will result in a match of the pressure dechne cun:e. .
loss from material balance connecting the end of injection with the start of shut-in. (We note that a similar spurt loss estimate for the KGD model IS given by
The second one matches the propagation pressure directly before the end of injection. V?5t~.5(bl + 1.478 x ml - pc)2
Both of these methods are more model dependent than the original Nolte method in Sp.mb = 0.771 J-t0.5'1.5h~5
the sense that they need the particular form of the width equation.
To illustrate the techniques, from now until the end of this chapter we will assume
'sf.L.5V?S(Pc - b) (9.119)
that Eqs. 9.40 and 9.41, i.e. the PKN width equation, hold during fracture propaga- + 1.02 h~Jt~.5(bl
-
+ 1.478 x ml - pc)
2'
tion.
(Material balance approach) At time t., when the pumps stop, there is an where the straight-line slope and intercept are derived from a g-plot generated with
instantaneous relocation of the fluid, but the average width should remain the a = ~.The similar estimate for a radial model is
same if we assume that the fracture does not propagate further. Thus, the material
balance provides an additional condition, i.e. the necessary equation to determine Vl/3 t2/3(b2 + 1.377 x m2 - Pc )8/3
spurt loss. Sp.mb=0.152 'e '2J-t2/3
220 Coupling of elasticity, flow and material balance Pressure decline analysis 221

1/3Vl/3( b) A useful byproduct of the technique is the estimate of the created radius
+ 1.23 1/3 fL i Pc - , (9.120)
t; (b2 + 1.377 x mZ - Pc )4/3 EIVl/3 1/3)
(9.126)
Rpr = 1.95 ( ;;/: (Ppr - PC)-4/3,
wher~ the straight-line slope and intercept are derived from a g-plot generated with
a = 9' A useful byproduct of the technique is the estimate of the created radius where the subscript "pr" indicates an estimate from the propagation pressure condi-
tion.)
EIVl/3j1.1/3) If the two approaches give similar results, i.e. the spurt-loss coefficients determined
Rmb = 1.45 ( ;;(3 (b2 + 1.377 x m2 - c )4/3, (9.121) from Eq. 9.118 and from Eq. 9.123 are almost the same, we have a consistent model
describing fracture propagation which can be used in design calculations. Since in
reality the fracture may propagate differently from what a model predicts, it cannot
where the subscript "mb" stands for material balance.)
be excluded that the spurt-loss coefficients determined from Eq. 9.118 and from
(Propagation pressure approach) If there is anything left we might be concerned
Eq. 9.123 are significantly different. Moreover, negative spurt-loss coefficient might
about, it is the observed pressure during the injection period. Indeed, the material
be also determined from the data. Any discrepancy means that the assumptions
balance technique provides a match for the pressure decline period, but there is no
underlying the method have to be revised.
guarantee that the simulated treatment pressure during the injection period will even
resemble the observed values.
To deal with this challenge, we have to characterize the pressure during propaga- Example 9.7 Pressure Decline Analysis, SFEJ Injection Test No.2
tion. In sharp contrast to the scenario after the pumps stop, the situation before the
The comparative study reported by Warpinski et al. [9] is based on a field experiment
pumps stop is rather steady. In other words the well bore treating pressure is almost named "OR! Staged Field Experiment No.3". By courtesy of Prof. S. A. Holditch and
constant and it is not difficult to pick a representative treating pressure directly before coworkers [23), we present here the data corresponding to the injection test No.2
the end of pumping. The pressure at the wellbore observed directly before the pumps carried out on March 16, 1989. Table 9.12 contains formation and fluid rheology data
stop is called the propagation pressure, Pw.pr. The propagation pressure approach for used for the analysis.
estimating spurt loss assumes that this value is known. In Table 9.13, we show the injection rate (notice that i corresponds to one wing) and
From Eqs. 9.40, and 9.43 the relation between the propagation pressure and frac- the calculated bottom hole pressure (obtained by adding a known hydrostatic pressure
ture length can be written as difference to a value measured at the surface). The time is relative to the start of
injection. The data were collected using a modem fracturing data acquisition system
(Crockett et al. [24]).
4 E '3fLV; ) =
Two items of Table 9.13 are set with italics numbers. Time t 37 min is the last
(Pw.pr - Pc) = 7.15 -4-- Xf (9.122)
(
h/te instant of the injection period and time t = 180 min is the assumed closure point.
Use the data of Tables 9.12 and 9.13 to determine the fluid-loss parameters and
Eliminating the unknown fracture length from Eq. 9.111 and 9.122, the alternative provide a consistent model of the fracture propagation, if possible. To simplify the data
estimate of the spurt-loss coefficient is treatment use the Newtonian fluid model with constant viscosity, estimated from the
power law parameters assuming a reasonable average width.
S _ 3.58E '3fLVi2 .
O.785hf(Pc - b)
p,pr - (
Pw.pr-pc)h/te
4 5 + E
I ' (9.123)
Solution

which assures the desired propagation pressure at the end of the injection. =
The end of the injection period is at t. 2220 s (37 min). The injected volume, V;,
(For the KGD model, the alternative spurt-loss estimate is given by is calculated as the time integral of the injection rate. The average injection rate is
=
i V;/te = 0.0564 m3/s.
1.02(pc - b)E'o.sfLo.5VO.5
Sp, pr = I +'0772(p pr _ C
P )zVO.5t05
I e (9.124)
Table 9_12 Formation and fluid data [9,23] for
se- - pc)2h~.5t~5 E'J.5h~5fL0.5 Example 9.7 .
' 6.13 X 1010 Pa 8.89 x 106 psi
and for the radial model by hf 51.8 m 170 ft
Pc 39.30 x 106 Pa 5700 psi
n 0.56
(9.125) K 0.622 Pa S.56 0.013 lbl- S0.56jft2
Pressure decline analysis 223
222 Coupling of elasticity, flow and material balance

Table 9.14 Injection test g-plot data for Example 9.7


Table 9.13 SFE3 Minifrac 2 data, March 16, 1989 [23] for Example 9.7
6/ 61D g(6ID, a) pw 6/ 6tD g(61D, a) pw
r i p", i I i
mlls MPa
t
min mlls
Pw
MPa min m'/s
Pk'
MPa min a = 4/5 MPa min a = 4/5 MPa
min
0.02703 1.460 43.54 73 1.973 3.175 40.38
0 0 27.61 38 0 43.54 114 0 40.32 1
0.08108 1.541 43.02 75 2.027 3.209 40.35
0.0069 27.61 40 0 43.02 116 0 40.29 3
1 42.72 77 2.081 3.242 40.32
118 0 40.27 5 0.1351 1.616
2 0.0230 41.33 42 0 42.72 3.276 40.29
7 0.1892 1.686 42.50 79 2.135
3 0.0263 41.17 44 0 42.50 120 0 40.24 40.27
9 0.2432 1.752 42.31 81 2.189 3.309
4 0.0339 41.41 46 0 42.31 122 0 40.21 40.24
11 0.2973 1.816 42.15 83 2.243 3.341
5 0.0386 41.62 48 0 42.15 124 0 40.19 40.21
13 0.3514 1.876 42.00 85 2.297 3.373
6 0.0386 41.71 50 0 42.00 126 0 40.16 40.19
15 0.4054 1.935 41.88 87 2.351 3.405
7 0.0452 42.06 52 0 41.88 128 0 40.14 40.16
17 0.4595 1.992 41.76 89 2.405 3.437
8 0.0511 42.25 54 0 41.76 130 0 40.12 40.14
19 0.5135 2.046 41.60 91 2.459 3.468
9 0.0488 42.22 56 0 41.60 132 0 40.10 40.12
21 0.5676 2.100 41.75 93 2.514 3.500
10 0.0519 42.33 58 0 41.75 134 0 40.06 40.10
23 0.6216 2.152 41.70 95 2.568 3.530
11 0.0577 42.58 60 0 41,70 136 0 40.04 40.06
25 0.6757 2.202 41.62 97 2.622 3.561
12 0.0637 43.03 62 0 41.62 138 0 40.02 40.04
27 0.7297 2.251 41.55 99 2.676 3.591
13 0.0642 43.19 64 0 41.55 140 0 40.00 40.02
29 0.7838 2.300 41.47 101 2.73 3.621
14 0.0642 43.33 66 0 41.47 142 0 39.97 40.00
31 0.8378 2.347 41.39 103 2.784 3.651
15 0.0641 43.48 68 0 41.39 144 0 39.95 39.97
33 0.8919 2.393 41.32 105 2.838 3.681
16 0.0640 43.58 70 0 41.32 146 0 39.93 39.95
35 0.9459 2.438 41.28 107 2.892 3.710
17 0.0642 43.71 72 0 41.28 148 0 39.90 39.93
37 1.000 2.483 41.21 109 2.946 3.739
18 0.0640 43.84 74 0 41.21 150 0 39.88 3.768 39.90
39 1.054 2.526 41.16 111 3.000
19 0.0640 43.98 76 0 41.16 152 0 39.87 3.797 39.88
41 1.108 2.569 41.10 113 3.054
20 0.0642 44.07 78 0 41.10 154 0 39.83 3.825 39.87
43 1.162 2.611 41.04 115 3.108
21 0.0642 44.14 80 0 41.04 156 0 39.82 3.853 39.83
45 1.216 2.652 40.98 117 3.162
22 0.0642 44.20 82 0 40.98 158 0 39.80 3.881 39.82
47 1.270 2.693 40.94 119 3.216
23 0.0642 44.26 84 0 40.94 160 0 39.79 3.909 39.80
49 1.324 2.733 40.86 121 3.270
24 0.0642 44.31 86 0 40.86 162 0 39.77 3.937 39.79
51 1.378 2.773 40.64 123 3.324
25 0.0640 44.37 88 0 40.64 164 0 39.56 3.964 39.77
53 1.432 2.812 40.73 125 3.378
26 0.0640 44.48 90 0 40.73 166 0 39.53 3.991 39.56
55 1.486 2.850 40.71 127 3.432
27 0.0640 44.51 92 0 40.71 168 0 39.47 4.018 39.53
57 1.541 2.888 40.68 129 3.486
28 0.0641 44.57 94 0 40.68 170 0 39.45 4.045 39.47
59 1.595 2.926 40.63 131 3.541
29 0.0640 44.60 96 0 40.63 172 0 39.41 4.072 39.45
61 1.649 2.962 40.59 133 3.595
3Q 0.0641 44.64 98 0 40.59 174 0 39.39 4.098 39.41
63 1.703 2.999 40.53 135 3.649
31 0.0642 44.67 100 0 40.53 176 0 39.37 4.125 39.39
65 1.757 3.035 40.42 137 3.703
32 0.0641 44.71 102 0 40.42 178 0 39.33 4.151 39.37
67 1.811 3.070 40.35 139 3.757
33 0.0635 44.73 104 0 40.35 180 0 3930 4.177 39.33
69 1.865 3.106 40.30 141 3.811
34 0.0636 44.76 106 0 40.30 182 0 39.27 4.203 39.30
71 1.919 3.140 40.41 143 3.865
35 0.0638 44.80 108 0 40.41 184 0 39.25
36 0.0635 44.82 110 0 40.38 186 0 39.23
37 0.0637 44.86 112 0 40.35 188 0 39.21 m = =
-1.208 x 106 Pa (-175 psi) and the intercept is b 4.432 X 107 Pa (6428 psi).
Note that the intercept is usually not on the plot, because it corresponds to g 0 and =
the actual data points have g values greater than 1.415. From Eq. 9.107, the leakoff
The shut-in period lasts from t =
38 min until the time the pressure decreases below
coefficient is
the closure pressure, t = 180 min. Table 9.14 shows the shut-in time, 6t, the corre-
sponding dimensionless shut-in time, 6tD, the g(6tD, a) values with ex = 3' and the C ----
-Jrhfm Jr X 51.8 x 1.208 x 106
wellbore pressure. L - 4.j;E' - 4..12220 x 6.13 x 1010
The plot of the wellbore bottomhole pressure vs. the g(6tD' ~) function is shown in
Figure 9.2. The straight line was determined by the method of least squares. The slope is
= 1.70 X 10-5 m/.Js = 4.32 X 10-4 ft/.rntiU.
_'--. ----_._-_ .. ,----" ---_. -------_._., '. ,--_.... ._-------_.,".._-_._ ._----,-- ._

224 Coupling of elasticity, flow and material balance Pressure decline analysis 225

~r---------------------------------, First we assume u. = 0.1 Pa- s (100 cp) and estimate the spurt loss from Eq. 9.118.
Substituting the known quantities
43

Sp.mb = 1.46 ( f-l,VfE'3)


--5--- (b + 1.415 x m - Pc)
-4
+
0.785hf(pc - b)
E'
hfte

= 1.46 (0.1 X 1252 x (6.l3 X 1010)3) [(44.32 _ 1.415 x 1.208 _ 39.3) x 106]:"'4
51.85 x 2220

40
+ 0.785 x 51.8 x (39.3 - 44.32) x 106 _ 0.00197 m (0.0775 in)
Pc=39.3 MPa 6.13 x 1010

1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 4 4.2 Using the PKN-a method, we simulate the injection test. For the shut-in period
g(AtD.4/5) Eq, 9.97 is used with the fracture area obtained at the end of injection. The calculated
pressures are shown in Figure 9.3 as curve a. Clearly, the calculated pressure follows
Figure 9.2 Straight-line fit of bottomhole pressure vs, g function the observed pressure decline with good accuracy. However, during the injection period
the simulated pressures are much higher than observed. The available data contradict
the hypothesis that the fracture propagated as a PKN fracture with Carter-type leakoff
Can we use the obtained leakoff coefficient without determining the spurt-loss mechanism and stopped propagating right at shut-in.
coefficient? The answer is no. Arbitrarily setting a spurt-loss coefficient equal to zero For comparison, the simulated pressure, assuming S p = 0, is also shown in the figure
would contradict Eq. 9.111 and distort the match of the shut-in pressure curve. as curve b. The zero spurt loss assumption does not correct the lack of match during
To use the results of the straight-line analysis correctly, we have to estimate the propagation and corrupts the fit of the decline period.
spurt-loss coefficient from the intercept of the straight line, b, using Eqs. 9.118 and The spurt loss coefficient can be adjusted to account for the fracture propagation
9.123. These equations involve the viscosity; hence we need a reasonable estimate pressure using Eq. 9.123. To use the equation, we need the propagation pressure, pw.pr
of the equivalent Newtonian viscosity of the fluid under average shear conditions. which is the wellbore pressure at t =
37 min and can be read from Table 9.14 to be
Table 9.15 shows equivalent Newtonian viscosities, calculated from Eq. 9.52 assuming pw.pr = 44.86 MPa. Substituting the known quantities into Eq. 9.123 yields
several average width values.

Table 9.15 Equivalent Newtonian viscosi-


ties for several average fracture
widths in Example 9.7
w f-l,.
m in Pas cp
0.002 0.0787 0.0297 29.7
0.004 0.158 0.0546 54.6
0.006 0.236 0.0780 78
0.008 0.315 0.100 100
0.010 0.394 0.122 122
0.012 0.472 0.144 144
0.014 0.551 0.164 164
0.016 0.630 0.185 185
0,018 0.709 0.205 205
0.020 0.787 0.225 225 o 50 100 150
0.022 0.866 0.245 245 TII'ne, t (min)
0.024 0.945 0.264 264
0.026 1.02 0.283 283 Figure 9.3 PKN-a simulation results for Example 9.7, IJ- = 0.1 Pa-s (100 cp); (a) Best
0.028 1.10 0.302 302 fit of the shut-in period, CL = 1.70 X 10-5 m/Sl/2 (4.32 X 10-4 ft/rninl/2) and
0.030 1.18 0.321 321 Sp = 0.00197 m (0.0775 in.); (b) Simulation with zero spurt-loss, CL 1.70 x =
10-5 m/SI/2 (4.32 X 10-4 ftJrninl/2) and Sp = 0
226 Coupling of elasticity, flow and material balance Pressure decHne analysis 227

357 x (6.13 x 1010)30.1 x 1252 0.785 x 51.8 x (39.3 - 4432) x 106 3.57 x (6.13 x 1010)30.3 x 1252 0,785 x 51.8 x (39.3 - 44.32) x 106
[(44.86 - 39.3) x 106J451.85 x 2220 + 6.13 x 1010
S -
P.P' - [(44.86 - 39.3) x 106]451.85 x 2220
+---------~--~--------
6.13 x 10 10

= -0.0017 m (-0.067 in). = 0.00155 m (0.0611 in).


Surprisingly, the spurt loss coefficient is negative. In other words, additional fluid
Both spurt-loss coefficients are positive, but different from each other. Figure 9.4
must enter the fracture to yield the observed propagation pressure. A negative spurt loss
shows that the material balance estimate provides an excellent fit of the pressure decline
is, of course, physically impossible and it indicates that one of the basic assumptions
period (curve a) while the propagation pressure estimate provides a satisfactory descrip-
inherent in the PKN model (i.e. "unretarded tip propagation") is not supported by the
observed data. tion of the injection period (curve b). The two estimates still contradict each other but
this contradiction can be easily resolved without rejecting the PKN model, as will be
In Chapter 10 we will provide a consistent interpretation of the SFE3 injection test
shown in the following. 0
assuming retarded tip propagation. However, in this chapter we deal with classical
models. For illustrative purposes, in the following we assume a much higher viscosity,
/L = 0.3 Pa s (300 cp). From Table 9.15 it is obvious that this average viscosity corre-
sponds to an unrealistic fracture width. Nevertheless, the high viscosity allows us to 9.6.4 Resolving Contradictions
show the technique to estimate fluid loss parameters in the cases where the classical
PKN model may be applied. The previous example showed the basic contradiction of pressure decline analysis:
=
Assuming /L 0.3 Pa s (300 cp) does not change the estimate of the leakoff coef- the inability to match propagation pressure and the pressure decline simultaneously.
ficient. The material balance estimate of the spurt loss coefficient is obtained from If a zero spurt loss is assumed, neither the propagation pressure nor the decline
Eq. 9.118 as 'part will be reproduced in general.
If the spurt loss is obtained from the material balance consideration, the decline
S _ 1 46 (0.3 x 1252 x (6.13 X 1010)3) 6-4
p,m/J - 51.85 x 2220 [(44.32 - 1.415 x 1.208 - 39.3) x 10 1 is described correctly but the propagation pressure may not be.
If the spurt loss is identified from the propagation pressure, the injection period
0.785 x 51.8 x (39.3 - 44.32) x 106 . will be matched correctly but the decline part will be shifted either up or down with
+ 6.13 X 1010 = 0.0126 m (0.495 10).
respect to the observations.
The propagation pressure estimate of the spurt loss coefficient is obtained from An additional degree of freedom is necessary to resolve the contradiction. Of
Eq. 9.123 as course, it is always possible to assume that the observed wellbore pressure is still
not the real propagation pressure. There might be a significant pressure drop through
4S the perforations. In addition, near the wellbore there might be a large pressure drop
induced by a tortuous flow path. If such a phenomenon can be identified and quan-
44
tified, then the additional pressure drop can be subtracted from the observations
<0 corresponding to the injection period.
ll. 43
::E For the SFE3 test and in most other cases, we have no evidence of significant
~
Q, 42 perforation or near wellbore pressure drop, and the introduction of an auxiliary
!
..
..,
::J

!
0..
41
pressure drop would mean that we practically discard the propagation pressure infor-
mation. For instance, in the case of the SFE3 injection test, one may attribute 1 MPa
(145 psi) to "tortuousity", but there is no real basis for doing this and 0.5 MPa
40 (73 psi) or 1.5 MPa (218 psi) could be also selected. We do not encourage this
c= 39.3M~a practice and suggest the use of more physically sound techniques.
39 One physically sound solution to the problem is to relax the strict postulate that
0 50 100 150 200
the fracture stops propagating exactly at the end of pumping. In this section, we
Time. t (min) investigate this extension of the theory.
Figure 9.4 Further simulation results for Example 9.7 and for Example 9.8, u. = The likely reason for overshooting the decline part of the pressure curve in
0.3 Pa-s (300 cp); (a) Material balance estiamte on the spurt loss, CL = Example 9.7, when the spurt loss is selected from the propagation pressure consid-
1.70 X 10-5 m/sl/2 (4.32 x 10-4 ftlminl/2) and Sp = 0.0126 m (0.495 in.); eration (i.e. curve b in Figure 9.4) is that the fracture may propagate for a while
(b) Propagation pressure estimate of the spurt-loss, CL = 1.70 x 10-5 rn/sl/2
after the pumps stop. Mathematically, the problem lies in the fact that the length
(4.32 X 10-4 ftlminl/2) and Sp = 0.00155 m (0.0611 in.); (c) Optimal three-
parameter fit, C L '" 1.70 x 10-5 rn/S1/2 (4.32 X 10-4 ft/min 1/2) and Sp = obtained during the propagation period and the spurt-loss coefficient (which was
0.000476 m (0.0611 in.), E.ta = 20 min used to obtain the length) do not satisfy the intercept condition of Eq. 9.111.
228 Coupling of elasticity, flow and material balance Pressure decline analysis 229

Allowing for a limited "after-growth" of the fracture after the pumps stop can not contradict the g-plot and should be limited to the interval where the pressure
correct the problem. The additional degree of freedom suggested in this section is points lie significantly above the straight line.
an incremental length growth, Sx], To simplify the derivations we assume that the Obviously, the leakoff coefficient is still estimated from Eq. 9.107. It should be
after-growth happens in a relatively short time period and the fitted straight line emphasized that the introduction of the after-growth can improve the fit only if the
already reflects the increased length. propagation pressure estimate of the spurt loss yields a simulated pressure decline
The average width instantly after shut-in, Wist, is a hypothetical width obtained curve which lies above the observations, since the incremental increase of the length
from the straight line directly after shut in. It is still given by Eg. 9.116 as decreases the calculated pressure in this period. However, if the propagation pressure
cannot be matched with a positive spurt loss or the relative position of the calculated
_ (b + 1.415 x m -
w- --------------~~~
pc);rhf
(9.127) pressure curves is the opposite, there is no way to obtain a consistent match within
lSI - 2E' .
the framework of the PKN model and retarded tip propagation should be assumed
This width corresponds to a length which is the propagation length plus the incre- as will be discussed in Chapter 10.
mental length growth. From volume balance,

(9.128) Example 9.8 Improved Pressure Decline Analysis of the SFE3 Injection
Test
where Wpr is the average width at the end of the injection period. At the end of
pumping Eq. 9.41 should hold, thus Continue the pressure decline analysis of Example 9.7 using the method of Section 9.6.4
=
but still assuming the artificially increased viscosity, M 0.3 Pa- s (300 cp).

X/.pr = (:~~r
(:~J, (9.129)
Solution
and W pr should correspond to the observed propagation pressure, prr (through The leakoff coefficient is exactly the same as in Example 9.7, i.e. CL 1.7 X =
Eqs, 9.41 and 9.43) according to 10-5 m/s1/2 (4.32 X 10-4 ft/minl/2). From Figure 9.7 it is seen that the propagation
pressure estimate of the spurt loss yields a calculated pressure decline curve lying
5E'Wpr
Pw,pr = Pc + ---.
2;rhf
(9.130) above the observations. Therefore, the introduction of a (preferably small) after-growth
can result in a consistent interpretation. The system of Eqs. 9.127 to 9.131 is written
substituting the known quantities as
Of course, the intercept must satisfy Eq. 9.111, but with the increased length which
is valid after the short "after-growth" period. Therefore, (44.32 - 1.415 x 1.208 - 39.3) x 106 x 1r x 51.8
WiST = 2 x 6.13 X 1010
_ Vj ;rhf(PC - b)
+ b...xf)hf + .
Sp- (9.131) Wist X (Xj.pr + 6.xf) + 2Spfuf = wprxf.pr,
2(xf,pr 4E'
The system of Eqs. 9.127 to 9.131 contains five unknowns, x/.PT> Sx], Sp, Wpr and xfpr -
_ (wpr)4
2.05
(6.13 X 1010 x 2220)
0.3 x 125 '
Wisi which are completely determined through the five equations. The solution of the
system can be found relatively easily, combining some algebraic Simplifications with 44.86 x 106 = 39.3 X 106 + 5 x 6.13 x 10108xwpr
a reliable root finding method. 2 x zr x 51.
The incremental length growth cannot be too large because "after-growth" may 125 tt x 51.8 x (39.3 - 44.32) x 106
happen only in a very limited time period. The time available for after-growth may s=
p 2 x (Xf.pr + !:"xf) x 51.8
+---__:_-:----:--::-:~--
4 x 6.13 X 1010
be well estimated from the g-plot as the interval where the observed pressures signif-
icantly deviate from the straight line. The estimated after-growth is realistic only if The solution of the system of equations is
this limited time is enough to reach the estimated b.x I: To check this the tip propa- Wpr = 0.00590 m (0.232 in.),
gation velocity has to be calculated. Once the spurt-loss coefficient is estimated, the
Wist = 0.00439 m (0.172 in.),
model can be solved in a "simulation mode". Using a small time difference directly
before the end of pumping, we can calculate another length, and the fracture prop- xf.pr = 247 m (813 ft),
agation velocity, uj, can be estimated from the length increment during the small
time step. Assuming a linearly decreasing propagation rate after the pumps stop, one
= 69.7 m (229 ft) and
!:"xf

can easily estimate the after-growth time from b.ta =


b.x f / (uJl2). This time should Sp = O.0tXl476 m (0.0187 in.).
__ __ ..
. ... ,-_. ---_., -----------------_-- ._-----

230 Coupling of elasticity, fJow and material balance Pressure decline analysis 231

To simulatethe whole pressurecurve we considerthree time periods.In the injection can be solved for wn. Once wn is known, qn is set equal to the value of either side.
period the PKN-a model is solved assuming the above-determinedleakoff and spurt The fracture length, x fe,' is constant and is obtained as the end value of the fracture
loss coefficients.In the after-growthperiod the first line of Eq. 9.105 is used, but with
length in the first cycle of calculations. (Of course, the after-growth concept can be
varying half-length. In a given time interval the increment of the fracture length is
introduced into this model as well.)
determined by the tip velocity. First the tip velocity equals the fracture propagation
rate determined numerically at the end of the injection period. Then this velocity is To start the second cycle of calculations, we need a starting value of Wn-l. which
decreased linearly to zero. The only additionalparameter entering the calculations is is the end value after the first cycle is accomplished. Also, all of the qj values are
the time of after-growth, /),to. "inherited" from the first cycle.
For the present example, if the fracture propagation is simulated with the esti- The outlined method is able to track the evolution of length and width during'
mated leakoff and spurt-loss coefficients,the tip propagation velocity is obtained as the injection period and the decrease of width during the shut-in period. The corre-
uf = 0.09 m/s (18 ft/min). Assuming linearly decreasing propagation rate (constant sponding net pressures are easily obtained from the width values using Eq. 9.89
deceleration),the after-growthperiod would last for more than 20 min. The long after- in the injection period and Pn.n = cfwn in the shut-in period. (Of course, if the
growth is not supportedby the g-plot but in this case we are less concernedbecause the observed propagation pressure cannot be matched with the simple Carter -type fluid
assumedviscosity is artificialand the calculationswere carried out only for illustration loss model, as was the case in Example 9.7, neither can the detailed leakoff model
of the technique. In spite of the artificial input data, the match of the pressure curve provide a match.)
is quite satisfactoryas seen from curve c in Figure 9.4. Note that for the simulation of
The exciting new feature of the detailed leakoff model is that it provides the
the pressure curve we use only three parameters estimated from the observed pressure
wellbore pressure even after fracture closure. To show this point, it is enough to
curve: the leakoff coefficient,the spurt loss and the length of the time interval after the
pumps stop, during which the tip propagationvelocity decreasesto zero. 0 realize that after fracture closure the actual leak-off rate, qn, is zero but the previous
non-zero injection history still causes a change of pressure according to Eq. 8.61. As
time elapses, more and more new leakoff rates will be zero and Eq. 8.61 provides a
9.6.5 Pressure Decline AnalYSis With Detailed Leakoff Description smooth return of the well bore pressure to the original reservoir pressure. This part
(Mayerhofer et al. Technique) of the pressure decline curve is an excellent source of information on the reservoir
permeability because the filter-cake resistance has no direct effect on these pressure
The basic method of Section 8.7 involving a detailed description of the flow in the points.
reservoir can be applied for pressure decline analysis as well. Ideally, the model Knowing the observed pressure decline curve, one may adjust one or more param-
should describe both the injection and shut-in periods. For brevity, we restrict our eters (first of all the final resistance of the filter-cake, Ro, and the reservoir perme-
consideration again to the PKN geometry. ability, kfb) by curve fitting. This is a "history matching" procedure. The technique
Assume that we know all the parameters governing the flow of fluid into and in of Mayerhofer and co-workers (25,26] differs from a simple history matching in one
the formation as listed in Section 8.7. In the injection period the model is solved as important respect. It accepts the observed pressures as given and inserts them into
described in Section 8.7. The result will be a history of flow rates and fracture lengths. the equation. Then Eq. 9.127 becomes
The final fracture length is denoted by xfe and the final fracture volume by Ve.
In the shut-in period, the same equations are used, except for the following minor hfxfn(Pobs,n-1 - Pobs.n)
changes. The injection rate is set to zero, and the length is' considered constant and cf6.t
set to the value xfe. Now the unknown variable is the width, and the net pressure
depends on the width according to p.; =
C iw, where the proportionality constant is
given by Eq, 9.100.
At time In (where In > to) we wish to determine wn. Since we know all of the
variables up to time tn-I, the following equation
hfxfe(wn-l-Wn) (9.133)
6.t
The only unknowns in Eq. 9.133 are Ro and kfb since the qn values have to satisfy
_ j.t [ n-l ~
(cfwn + Pc - Pr)--k h -Qn-IPD(tn - tn-l)+:L (qj - qj-l)PO(tn - tj-l) qn = hjXfn(Pobs,n-l - Pobs,n) (9.134)
7r jb f j=1
cf6.t
for every tn > to and hence are given by the left hand sides of the already satisfied
equations. (The reservoir pressure, P" the fracturing fluid viscosity, u.f' the reservoir
(9.132) fluid viscosity, u, and the fracture height, hj, are assumed to be known.)
References 233
232 Coupling of elasticity, flow and material balance

11. Lenoach, B.: HydraulicFracture ModellingBased on AnalyticalNear-TipSolutions, in


The dimensionless pressures, PD, correspond to solutions of transient flow through
Computer Methods and Advances in Geomechanics, Siriwardane, H.J. and Zaman, M.M.
porous media such as Gringarten and Ramey's [27] infinite conductivity fracture
(eds.), AA Balkerna, Rotterdam, 1994.
solution or Cinco-Ley and Meng's [28] solution for a naturally fractured reservoir 12. Nolte, K.G.: Determinationof Fracture Parametersfrom Fracturing Pressure Decline,
with a finite conductivity fracture. These solutions are available in several forms such Paper SPE 8341 presentedat the SPE AnnualTechnicalConferenceand Exhibition,Las
as tables, algorithms, or curves. The pressure decline analysis module computes the Vegas,Sept. 23-26, 1979.
dimensionless time differences tn - tn-l, tn-l - tn-2, ... using the actual fracture 13. Nolte, KG.: Fracturing-PressureAnalysis,in Recent Advances in Hydraulic Fracturing,
length (corresponding to time tn) and passes them to the module for calculating PD. Gidley, J et al. (eds.)., Monograph Series, SPE, Richardson, Texas, Vol. 12, SPE,
The module can be selected according to our knowledge of the actual structure of Richardson,TX, 1989.
the formation. Since the fracture length at time tn is known only at time tn, the 14. Nolte, KG.: Determinationof Proppant and Fluid Schedules from Fracturing Pressure
calculation of the dimensionless pressures has to be done again and again in every Decline, SPEPE, (July), 255-265, 1986.
15. Nolte, KG.: FractureDesignConsiderationsBasedon PressureAnalysis,SPEPE, (Feb.),
time step during the propagation period, but it can be done for the rest of the period
23-30, 1988.
only once if the final fracture length is already known.
16. Nolte, K.G., Mack, M.G. and Lie W.L.: A SystematicMethod for Applying Fracturing
The obtained system of equations allows the best estimate of the value of Ro and Pressure Decline: Part 1., Paper SPE 25845 presented at the SPE Rocky Mountain
k fb through a least squares fit. Regional/LowPermeabilityReservoirs Symposium,Denver, CO, April 12-14, 1993.
The technique of Mayerhofer and co-workers employs standard methodologies of 17. Castillo,J.L.: ModifiedPressure Decline Analysis IncludingPressure-DependentLeak-
modem pressure transient analysis, including the log-log plot diagnostic depiction of off, paper SPE 16417presented at the SPE RockyMountain Regional/LowPermeability
the rate-normalized pressure and its derivative. Reservoirs Symposium,Denver, CO, May 18-19, 1987.
Our presentation of both the Nolte and the Mayerhofer techniques considers only 18. Zhu, D and Hill AD.: The Effect of Temperatureon Minifrac Pressure Decline, Paper
the basics of the methods. The interested reader may find several extensions, varia- SPE 22874, 1991.
tions and considerations in the original papers [12-17,22,25,26]. 19. Valk6, P. and Economides,MJ.: Fracture HeightContainmentWithContinuumDamage
Mechanics,Paper SPE 26598, 1993.
20. Abramovitz,M. and Stegun, LA (ed.): Handbook of Mathematical Functions, Dover,
References New York, 1972.
21. Wolfram,S.: Mathematica: A System for Doing Mathematics by Computer, 2nd ed.,
Addison-Wesley,NY, 1991.
1. Perkins, T.K and Kern, L.R.: Width of Hydraulic Fractures,1PT, (Sept.), 937-949, 22. ShlyapoberskyJ., Walhaug,W.W., Sheffield.R.E., Huckabee,P.T.: Field Dete.remina-
1961; Trans. AIME, 222. tion of Fracturing Parameters for Overpressure Calibrated Design of Hydraulic Frac-
2. Khristianovitch,S.A and Zheltov, y'P.: Formation of Vertical Fractures by Means of turing, Paper SPE 18195, 63rd Annual Technical Conference and Exh. Houston, TX,
Highly ViscousFluids, Proc. World Pet. Cong., Rome, 2,579, 1955. October 2-5, 1988.
3. Zheltov,Y.P. and Khristianovitch,S.A: On the Mechanismof Hydraulic Fracture of an 23. Holditch, SA.: Personal Communication, 1993.
Oil-bearingStratum, Izv. AN SSSR, OTN, (No 5), 3-41, 1955. 24. Crockett,AR., Okusu,N.M. and Cleary, M.P.: A Complete Integrated Model for
4. Geertsma,J. and de Klerk, F.: A Rapid Method of Predicting Width and Extent of Design and Real-Time Analysis of Hydraulic Fracture Operations, Paper SPE 15069,
HydraulicallyInduced Fractures,lPT, (Dec.), 1571-1581, 1969. presentedat the 56th California Regional Meeting, Oakland, CA, 1986. .
5. Geertsma,J.: Two-dimensionalFracture-PropagationModels, in Recent Advances in 25. Mayerhofer,M.I., Economides,M.J. and Ehlig-Economides,CA.: Pressure Transient
Hydraulic Fracturing, Gidley, J.L. et at. (eds.), SPE Monograph 12, SPE, Richardson, Analysis of Fracture Calibration Tests, Paper SPE 26527 presented at 68th Annual
TIC, 1989. TechnicalConferenceand Exhibition, Houston, Texas,3-6 October, 1993.
6. Nordgren, R.P.: Propagation of a Vertical Hydraulic Fracture, SPEJ, 306-314, (Aug.) 26. Ehlig-Economides,CA., Fan Y. and Economides,M.J.: InterpretationModel for Frac-
1972; Trans. AlME, 253. ture CalibrationTests in Naturally Fractured Reservoirs, Paper SPE 28690 presented at
7. Economides,M.1., Hill, AD. and Enlig-Economides,C.A: Petroleum Production the InternationalPetroleumConference & Exhibition,Veracruz,Mexico,October 10-13,
Systems, Prentice Hall, Englewood Cliffs, NJ., 1994. 1994.
8. Nierode, D.E. Simple Calculation of Fracture Dimensions, in Recent Advances in 27. Gringarten,AC. and Ramey, AJ., Jr.: Unsteady State Pressure Distributions Created
Hydraulic Fracturing, Gidley, J.L. et al. (eds.), SPE Monograph 12, SPE, Richardson, by a Wellwith a Single-InfiniteConductivity VerticalFracture, SPEJ, (Aug.), 347-360,
TX,1989. 1974.
9. WarpinskiN.R., Moschovidis,Z.A, Parker C.D. and Abou-Sayed,I.S.: Comparison 28. Cinco-Ley,H. and Meng, H.Z.: Pressure Transient Analysis' of Wells with Finite
Study of HydraulicFracturingModels:Test Case - GRI-StagedField Experiment No.3, Conductivity Vertical Fracture in Double Porosity Reservoirs, Paper SPE 18172
SPE Production & Facilities, 9(1), 7-16, 1994. presented at the 63rd Annual Technical Conference and Exhibition of the SPE.held
10. Kemp, L.F.: Study of Nordgren's Equation of Hydraulic Fracturing, SPE Production in Houston,TX, October 2-5, 1988.
Engineering, (Aug.), 311-314, 1990.
10
FRACTURE PROPAGATION

In Chapter 9 we presented a vision of a fracture resulting from the combination of a


fundamental law (material balance) and constitutive equations which describe elas-
ticity and fluid rheology. The models were intended to be representations of certain
physical processes having major influence on the final geometry of the created frac-
ture. However, in the case of hydraulic fracturing, these models present a rather sterile
view. Stating explicitly, assuming implicitly, or sometimes even unconsciously, the
fracture propagation velocity with which the tip propagates away from the well is
often considered to be a trivial consequence of the other processes described in a
model. Of course, more realistic fracture propagation modeling requires the under-
standing of the processes involved at the tip.
The fracture propagation velocity is a crucial issue in predicting not only the
instantaneous geometry but also the behavior of the propagating fracture which, in
tum, has a compounding effect on the resulting geometry.
What should the purpose of propagation modeling be? If the purpose is to repro-
duce a specific observed behavior, then two problems may arise. First, the observed
behavior (such as the fracture treatment pressure) may constitute only a small window
in a far more complicated global process. A model that can match such a window
may fail considerably in accounting for unobserved behavior before or after the
specific observation. Second, more than one model may reproduce the same data.
This can be done by emphasizing or even arbitrarily weighing different factors which
may prove wrong outside the window of observation or even .be incompatible with
other physical phenomena.
The process of selecting a model or parameters that can both reproduce observed
behavior and conform to other information about the system, while respecting funda-
mental laws and/or common sense, is the essential element of data interpretation.
Modelers should never become overconfident just because they are able to repro-
duce one set of observations or a particular presentation style for acquired data
(especially if they achieve this by adjusting several parameters). The significant
contribution from a model is the identification of characteristic behavior trends in its
forward simulations that are distinct from the behavior of other models and whether
these trends have been observed in acquired data. The applied value of the model is
further enhanced with analyses demonstrating the sensitivity of the model behavior
to parameter values that are important for practical decision-making.
236 Fracture propagation Fracture mechanics 237

It would be useful if the developers of hydraulic fracture models were to adopt a theories. At issue is the velocity with which the fracture tips (lateral and vertical)
common method of presentation with which to distinguish the various models and the travel, the causes of possible retardation (with respect to the one obtained from
effects of various phenomena. Petroleum engineers have adopted the log-log plot of some trivial assumptions) and the resulting observable net fracturing pressure vs.
pressure change and its derivative with respect to the log of elapsed time as a standard time behavior.
for presentation of pressure transient data acquired at a constant flow rate: Variations Clearly, it is this debate that could benefit from some common language and
on this plot for data acquired with a constant wellhead pressure or for variable uniformity in the representation of results.
pressure and rate conditions have also been developed. The derivative presentation In the following, we give a brief overview of fracture mechanics concepts relevant
has proved to be highly effective in its sensitivity to model and parameter variations to tip propagation and describe their usage in hydraulic fracture modeling.
while, at the same time, these presentations reflect simple straight-line derivative
trends with which behavior of practical interest can be readily identified. With the
exception of the qualitative Nolte-Smith [1] analysis for the simple 2D models, no 10.1 Fracture Mechanics
serious attempt has been made among hydraulic fracturing workers in this direction.
The scientific-engineering discipline of fracture mechanics has evolved from the
There are three basic requirements in constructing a fracture propagation model:
necessity to avoid the dangerous collapse of engineering structures such as bridges
L Fundamental laws such as material and energy balances must be obeyed in a or dams. With loading tests of carefully selected specimens, the limits of safe loading
demonstrable way. can be determined experimentally. After experimental data are collected, one can
2. A complete mathematical formulation of the governing and boundary equations, attempt to formulate rules to avoid failure. In formulating a failure criterion, concepts
of the theory of elasticity are used (e.g. deformation, stress, stress intensity factor),
without resorting to arbitrary "weighing factors", should be derived. Any simpli-
but the criterion itself does not follow from the theory of elasticity alone. As a matter
fying assumptions must be defined clearly and their effects quantified. The
of fact, the process of failure or fracture lies outside the area of elasticity.
formulation of the model has to be separated from its computer algorithm.
For a long time the theory of fracture mechanics has been biased by its ultimate
3. A fracture tip propagation criterion and its interaction with the provided energy aim, i.e, to avoid fracture. Von Karman [2] realized that there are two schools of
must be explicitly stated. thought in fracture mechanics: one aimed to provide simply definable conditions to
remain on the safe side and the other more concerned about the physical image of the
A potentially very powerful tool is the formulation of analytical solutions with
evolution and propagation of the fracture once the loading limit has been exceeded.
explicit simplifying assumptions in the governing equations. These solutions should
While the overwhelming majority of works in fracture mechanics corresponds to the
serve as limiting forms of the same governing formulations without the simplifying
first group, the second group is of major interest to hydraulic fracturing where the
assumptions. The analytical forms may suggest presentations of the numerical simu-
aim is to create a fracture and not to avoid it.
lations (such as the log-log plot of pressure change and its derivative) that may help
The history of fracture mechanics is described in excellent books by Timo-
in the intuitive understanding of the implications of the non-simplified cases. A suit-
shenko [3] and Gramberg (4). The first and up to now the most influential ideas
able presentation also provides a means to demonstrate the sensitivity of the model
go back to Coulomb [5] and Mohr [6], who assumed that the collapse of the mate-
behavior to the parameters eliminated from the analytical solution.
rial takes place when the state of stress of a material point reaches the boundary
A common methodology of representation, emanating from the analytical solution
of the safe domain. They assumed that the boundary can be described in simple
(e.g., net pressure during fracturing on log-log coordinates) should be adopted also for
algebraic terms and considered the location of the boundary as a material property.
numerical modeling, not only to provide a recognizable pattern but also to indicate the
Some selected points on this boundary are referred to as the strength of the
extent and manner of deviation from the simplifying assumptions. Many numerical material. The tensile strength is the limiting value of stress the material can bear in a
simulations of the same observed behavior which may look credible in a Cartesian uniaxial test where the loading is tensile. The compressive strength is another value
representation, for example, may fail miserably when subjected to the rigors and (usually an order of magnitude higher) which has to be applied to cause failure if in
scrutiny expected from an analytical solution. More to the point, the use of analytical the uniaxial test the stress is compressive.
solutions should be preferred for cases in which aspects that would require use of While the Coulomb-Mohr theory provides a framework to describe the strength
the numerical model are insignificant. of materials with respect to the geometry of loading conditions, it was Griffith [7,8]
Fracture propagation is an area that has spun considerable debate and controversy. who first gave a theory to explain some aspects of the strength behavior. He was
While in Chapters 1 to 9 of this book we introduced several new developments, and concerned with "brittle" materials, first of all glass. Jaeger and Cook [9], looking
while several issues still need investigation, the means of coupling material balance, back to the history of fracture mechanics, gave a very objective appreciation of
rock elasticity and fluid rheology is reasonably well understood and accepted. The Griffith's contribution as follows: "His concept of failure, based on the existence
mechanism of fracture propagation, on the other hand, is reflected in several different of minute internal and surface flaws, is questionable to rock. However, the body of
238 Fracture propagation Fracture mechanics 239

theory which has grown up around his concept is very useful. .. and there seems to is easily derived from Eq. 2.68. The result is
be no reason why it should not be applied to the behavior of much larger cracks ... "
~w - ~V _ 21(pooc~C 21(P68c~c
(10.4)
p - Po - Po E' E'
10.1.1 Griffith's Analysis of Crack Stability Comparing Eqs. 10.3 and 10.4 Griffith [7,8] realized that'half of the work done by
th~ inner pressure, i.e, ~ Wp/2, is not converted into strain energy. What happens
Let us consider a slice of thickness 8 from an infinite material containing a three-
WIth the other half of the work? Part of it is used to create new surface and if
dimensional extension of a "line crack" of half-length c, as depicted in Figure 10.1.
still any surplus energy is available, it is dissipated as heat. Whatever the' energy
The situation is described well by a plane strain condition. We assume that the crack
consumption is during the creation of the new surface, in a given material it should
is opened by a constant pressure, po, both with respect to location and time. The
depend only on the area of the created new surface. Griffith postulated that there
strain energy stored in one half of the medium (i.e. corresponding to one wing of
is a material property, the specific surface energy, a, that characterizes the energy
the two-wing crack) is calculated as (see Eq. 2.64)
consumption while a unit area of new surface is created. The specific surface energy
is measured in J/m2
W = JrP58? (10.1) Once we accept that a is a property of the material, an energy balance can be
o 2" written as
If the crack propagates, a next state (also depicted in Figure 10.1) will be character- ~ W p = ~Wo + 2ct86.c + generated heat. (10.5)
ized by the strain energy: The crack will not propagate if the left-hand side is not enough to cover the two first
items on the right-hand side. This is the famous Griffith stability criterion, stating
W + ~W = Jrp6(c + ~d (10.2) that a crack is stable if
a 0 2"
~Wp:::; ~Wo+2aMc. (10.6)
from which the increase of strain energy is obtained (neglecting the second-order After substitution of Eqs. 10.3 and 10.4 into inequality 10.6, the inequality becomes
term) as
2JrP60c~c JrP50c~c
(10.3) E' :::; E' + 2a8~c, (10.7)

The work done by the inner pressure while moving the fracture faces apart and The inequality can be rearranged such that the variables characterizing a given state
ahead is obtained as pressure multiplied by volume change, where the volume change of the material are on the left and variables, characterizing the material behavior are
on the right-hand side. Then, the inequality becomes

2 2aE'
poc:::; --, (10.8)
n
i.e,
Strain energy stored 2aE')1/2
poc1/2:::; ( -- (10.9)
n

It is not difficult to discover the stress intensity factor, K/ = pocl/2, on the left-
hand side (see Eq. 2.53). Thus, the Griffith inequality can be interpreted in terms of
stress intensity. There exists a critical value of the stress intensity factor, often called
f.+.----- c fracture toughness, KIC, with the following property: If the stress intensity factor at
the tip is less than the fracture toughness, the fracture is stable. Otherwise, the fracture
is unstable, and starts to propagate. A propagation process through equilibrium states
would imply that the equality
(10.10)

Figure 10.1 Propagation of a line crack holds during the whole process.
240 Fracture propagation Fracture mechanics 241

From inequality 10.9 the fracture toughness is obtained as 10.1.2 Mott's Theory for the Rate of Crack Growth

K/c
2a/)
= ( --n
1/2
(10.11)
When a crack propagates, the velocity of certain material points has to be altered.
Griffith's criterion is not sufficient to describe the situation because of its static
nature. Matt [14] presented a simple theory of the crack propagation velocity taking
Unfortunately, Eq. 10.11 is not very useful to determine the fracture toughness, account of the kinetic energy change. He augmented the energy balance, i.e. Eq. 10.5,
because a is not an easily observable or derivable quantity. On the other hand, it is introducing the change of kinetic energy, LlW k. but neglecting the irreversible heat
not difficult to accept that the stress intensity factor (even if it is connected with an generation term:
otherwise infinite stress state and even if it has a rather unusual physical dimension) LlW p =LlW0 + 2ao!:.c + 11WK. (10.14)
is a state variable and that the material responds to it by rupture if the value is above Mott gave a gross estimate of the kinetic energy for a crack increasing its half length
a certain limit. and width. If u(x, y) is the displacement at any point corresponding to a half length
The fracture toughness of a given material can be determined from experiments c, the velocity at any point is the first derivative of the displacement with respect
on specimens of specific shape, observing the critical loading conditions when the to c multiplied by the tip velocity, dc/dr, Once the velocity is known, the kinetic
fracture starts to propagate. It is not necessary to establish a given stress intensity energy can be calculated according to
factor by pressurizing a crack; the same effect can be achieved by applying a tensile
stress, 0"0, because then K/ = O"OCI/2 at the tip.
I (dC) 2 JOOJC (dU)2 (10.15)
, Griffith's concept has given birth to the scientific and engineering discipline called W K = 'i_op dt dc dxdy,
linear elastic fracture mechanics, LEFM. The discipline provides an efficient frame- cc 0
work to measure, distill, store and reveal information on the rupture properties of where p is the density. Mott did not carry out a rigorous calculation of the integral.
materials. It is particularly useful in predicting critical loading conditions for struc- Instead, he used an order of magnitude estimate. From Eq. 2.36,
tures under design and in use.
The success of LEFM lies in the fact that, for certain materials and for a certain du Po
(10.16)
scale domain, the load-carrying capability of the material containing a crack is dc ::::::1'
inversely proportional to the square root of the crack length. The longer the crack, at least near the fracture surface, i.e. in an area approximately c2 large. Thus,
the easier it is to propagate. Eq. 10.15 can be rewritten as
An analogous derivation for a radial crack results in the condition of stability
dC)2 P6c2
(10.12) W K = kop ( dt '2' (10.17)

where k is a numerical factor with unit order of magnitude. The change in the kinetic
which can be rewritten using Eq. 10.11 as
energy during a time interval Llt is

+ (dC)2]
2 1/2 d2c
- poR
it
..s K/c (10.13) 2kopp~c [cd(2 dt LlC

Again, for a stable crack the product of the (net) pressure and the square root of the '2
(10.18)
characteristic dimension (now radius) has to be below a certain critical value.
where we have substituted 11c = (dcfdt)Lll. Combining Eqs. 10.3, 10.4, 10.14 and
If this inverse proportionality can be observed in practice, there is no reason
10.18 gives the energy balance in the form
to preclude LEFM. However, for more sophisticated problems, including effects of
large-scale yielding, time dependence, non-trivial loading conditions and effects of
damage, the inverse square root proportionality Simply does not hold and it is difficult
to use classical fracture mechanics concepts (Lamaitre [10]). The train of thought (10.19)
applicable under these "extreme" conditions is termed local approach. Very often
other names-process zone theory (Boone et al. [11]), continuum damage mechanics i.e.
(Kachanov [12]) and crack layer' theory (Chudnovsky [13])-are used to indicate devi- (10.20)
ations from LEFM.
242 Fracture propagation Classical crack propagation criterion 243

Mott assumed that the crack propagates from an equilibrium state, so the initial value in such a way that the stress intensity factor, K[, is kept nearly equal to the critical
of c can be expressed from inequality 10.8, used as an equality: stress intensity factor, K[c, during crack extension at each node (Clifton and Abou-
Sayed [22]) or, in other words, "growth occurs when K[ reaches K[c" (Thiercelin
2E'a et at [23]). The rebuilding of the critical loading conditions is accomplished through
co=-- (10.21)
:rcp~ fluid flow, which is described in different levels of complexity from one-dimensional
steady-state flow to two-dimensional unsteady flow involving a possible non-wetted
Also, he assumed that dc/dt and d2c/dt2 are zero at the start of propagation. The solu- zone at the tip.
tion of Eq. 10.20, assuming the above initial conditions, was found by Mott [14] as Unfortunately, the direct application of classical fracture mechanics concepts in
hydraulic fracturing has not been as successful as in many other fields, e.g. in the
dc = r;E (1_ Co) . (10.22) analysis of concrete structures.
dt V 2kP C

According to the theory, once the crack starts to propagate, it rapidly accelerates, 10.2.2 The Injection Rate Dependence Paradox
and at later times, when the length is much larger than the initial length, the prop-
agation rate is almost constant. Experiments show that for brittle solids, a limiting To illustrate why the classical crack propagation criterion has proved less than
velocity is often about half the shear wave speed (Billington and Tate [15]). Clearly, successful in hydraulic fracturing, we consider a somewhat simplified but illustrative
Mott's theory provides only an upper estimate on the propagation rate since the example.
energy dissipation due to irreversible processes is neglected.

Example 10.1 Radial Fracture Propagating According to LEFM


10.2 Classical Crack Propagation Criterion for =
Calculate the net pressure at t 200 min in a radially propagating fracture for the plane
Hydraulic Fracturing strain modulus, fracture toughness and injection rate given in Table 10.1. Assume that
there is no leak-off, the viscosity of the fluid is so low that the pressure drop in the
10.2.1 Fracture Toughness Criterion fracture is negligible and that the fracture propagates according to the LEFM criterion,
i.e. at every time instant the radius R satisfies inequality 10.13 "sharply", with the
Based on the original ideas of Griffith, fracture propagation is often modeled by inequality sign becoming an equal sign. Finally, repeat the calculation of net pressure
applying the following cycle of steps: (1) calculate the stress intensity factors for assuming an injection rate that is ten times smaller.
a given geometry and loading condition, (2) apply the stability criterion involving
fracture toughness to determine whether the fracture will propagate, and (3) if the
fracture is unstable, propagate it a certain distance corresponding to the selected Solution
time step.
Since there is no leakoff, the volume of one wing (see Eq. 2.30) equals the injected
The first step is accomplished by the perhaps most successful scientific-engineering volume:
method of this century, the finite element method (Zienkiewicz [16]) or by one of its 8R3 PnO .
variations. The second step is the application of inequality 10.9, or one of its many --=It.
3'
generalizations (Erdogan and Sih [17]; Sib [18]). The third step is often arbitrary,
and sometimes hidden in the computer program, and can be as simple as "add one On the other hand, since the fracture propagates according to LEFM, from
element per time step". The tip propagation velocity is obtained implicitly, through "equation" 10.13
IjZ it
the rate at which the loading conditions build back up to instability. The numer- poR = -K/c
2
ical models often use interactive computer graphics (Boone et al. [19]) and adaptive
remeshing (Gerstle and Xie [20]).
In hydraulic fracturing, the geometry and loading conditions can be assumed to Table 10.1 Input data for Example 10.1
be rather simple, e.g. an ellipse may be assumed (Bouteca [21]) because detailed 6.13 x 1010 Pa 8.89 X 106 psi
information on the peculiarities of the shape is not available. The stress intensity 7.6 x 106 Pa. mI/2 2000 psi- ftI/2
factor at the tip, KJ, is calculated from the actual pressure distribution inside the 0.0662 m3Js 50/2 = 25 bpm
fracture. The propagation criterion is usually the simplest one: "crack advance occurs
244 Fracture propagation Retarded fracture propagation 245

Eliminating the unknown R from the above two equations we obtain 10.3 Retarded Fracture Propagation

p".o
= (T(6)
24
1/5 (KYc)
E'i
1/5 (~)
t
1/5 There is continuously growing evidence that existing fracture propagation models
neglect some important aspects of the near tip region. Often, the observed field
treating pressure is much higher than predicted by these models (Medlin and
(7.6 X 106)6 ] 1/5 ( 1 ) 1/5
=2.09 x
[ 6.1 x 1010x 0.0662 -- Fitch [26]; Palmer and Veatch [27]). The net pressure is rather insensitive to limited
12000
rate variations and especially to fluid viscosity changes (Cleary [28]).
= 0.69 MPa (100 psi). Several efforts have been made to build more realistic models. The difficulty lies
in the fact that fracture propagation is a result of several mechanisms, i.e. a model
If we repeat the calculation with the much smaller injection rate, i 0.0066 m3/s = contains several elements. If one of these elements is modified, the performance can
(2i = 5 bpm), then the net pressure will be higher, i.e. 1.1 MPa (159 psi), and this
result is difficult to accept. 0 be predicted only using a fully coupled implementation of the new modeL Often the
results are disappointing because the introduced modification does not generate the
In laboratory experiments the decreasing pressure vs. time curve can be observed
desired change in overall behavior. In the following we give a brief description of
clearly (Heuze et al. [24]; De Pater et al. [25]), and this is often considered as proof
some current trends in modeling propagation of hydraulically induced fractures.
of the LEFM theory. However, the result that the net pressure is inversely propor-
tional to the ! power of the injection rate has never been confirmed in laboratory
experiments and obviously contradicts the everyday field experience of fracturing 10.3.1 Fluid Lag
workers.
The LEFM theory gives a false picture of the fracturing process in large scale and In order to overcome the paradoxical predictions of LEFM, the concept of fluid lag
under compressive far field stress states. What is the reason for this discrepancy? (non-wetted zone, non-penetrated zone) has been reintroduced (Jeffrey [29]). The
In large scale hydraulic fracturing, Eq, 10.5 still holds (with the modification to concept was first introduced in the literature by Khristianovitch and Zheltov [30].
include the kinetic energy change, as suggested by Matt). The dissipated energy in The additional degree of freedom allows one to "tune" the pressure without giving
the form of heat is, however, no longer a negligible term. It becomes more and more up Eq. 10.10.
important with size. The existence of a large fracture generates a zone of microcracks Is there a fluid lag in reality? Of course there is. It occurs because fracture prop-
around and ahead of the fracture. This zone can be called damaged zone, cohesive agation is a discrete process with "jumps". The exact location of the fluid front
zone, process zone, crack layer or crack band. At large scales more and more energy depends on several factors, e.g. when the last jump occurred and the microscopic
is used up by this process and by possible other irreversible changes generated in shape of the flow channel. Additional factors such as interfacial tension and the
the surrounding formation. Thus, with increasing scale the predictions of LEFM, wetting angle might be of primary importance when formulating a moving boundary
neglecting the dissipation term, become more and more unrealistic and contradict condition for the fluid. Without a clearly stated boundary condition the COnceptdoes
common sense. not enlighten the mechanism controlling tip propagation.
When viscous forces become important, an additional dissipation of energy occurs We agree with Shlyapobersky and Chudnovsky [31) who state that "... often these
due to fluid flow. This part of the energy dissipation, i.e. the frictional pressure drop, models, however, simulate too high net pressures at early injection time and too low
is incorporated into almost all existing models as was shown in Chapter 9. net pressures for a large, poorly contained fracture at later time ... The inability of
It is not difficult to see that classical fracture mechanics, coupled with mate- these numerically very accurate hydraulic fracturing models to simulate high net
rial balance, elasticity and fluid flow description, gives an exaggerated estimate of pressures calls for modifying these models by including additional effects that may
the tip propagation rate, especially at larger injection times. The two-dimensional be left unaccounted for. .. "
models of Chapter 9, namely PKN, KGD, radial and especially the Nordgren-Kemp
formulation estimate the tip velocity more realistically. In these models, the fracture 10.3.2 Tip Dilatancy
propagation rate is determined as the limiting velocity of the fluid approaching the
tip, calculated as if the fluid were flowing in a static ("frozen") counterpart of the Plastic rock behavior near the tip may cause rock dilatancy, constrain the opening
instantaneous geometry. In other words, no additional dissipation of energy near the and lead to increased resistance of fluid flow [32-34]. Shlyapobersky and Chud-
tip region is considered. Technically, this is achieved by explicitly (or unconsciously) novsky [31] find that " ... tip dilatancy can only affect the net pressure if the charac-
setting the net pressure to zero at the tip. An analogy might be a knife cutting butter, teristics of non-linear deformation and failure of rock mass are subject to a strong
where the velocity of the knife is controlled by the friction between the surface of scale effect, which may be difficult to predict based on laboratory data."
the knife and the butter but not by the processes at the edge. In the following we The idea of a region with a "narrower" flow channel near the tip (i.e. a stiffer
refer to the "knife in the butter" analogy as unretarded propagation rate. region with higher elastic modulus) has been investigated using a fully coupled model
246 Fracture propagation Continuum damage mechanics 247

(Gardner [35J). It seems that such modification cannot reproduce the increasing char- extreme cases are considered. In the "cooperative" fracture case, the intensity of the
acter of the net pressure curves at late injection times. damage formed as a response to the stress concentration at the tip of a propagating
There is growing evidence that any attempt to vary numerical values of the param- crack is much greater than the intensity of pre-existing defect population. This case
eters involved in the classical models (e.g. elastic modulus), especially if done only is well described by the crack layer theory [13]. The other extreme mode is when
in the near tip region, is insufficient to achieve a breakthrough in understanding the crack propagation is controlled by the field of pre-existing defects and does not cause
propagation process. noticeable changes in the field. This case is described by statistical fracture mechanics
(Chudnovsky and Gorelik [40]). Both mechanisms are present in hydraulic fracturing
10.3.3 Apparent Fracture Toughness and are subject to scale effects. The translation of these creative concepts into working
models of hydraulic fracturing might be a major improvement in the future.
This approach has emerged from the observation that observed fracture toughness is
not a material property (Shlyapobersky et al. [36]). The increase of apparent fracture
toughness with high confining stress can be predicted by theory (Yew and Liu [37]). 10.3.5 The Reopening Paradox
Apparent fracture toughness is widely used in the industry as a calibration parameter
in order to reproduce elevated net pressures. The idea is to reproduce the high net Before the main hydraulic fracturing treatment, an injection test with the same fluid
pressure in the injection period either exactly the same way as our Example 10.1 and the same injection rate but smaller injected volume is often executed to reveal
shows or to postulate that the net propagation pressure is a sum of the unretarded the Ieakoff characteristics of the formation. Thus, during the main treatment - or
propagation pressure plus the propagation pressure increase due to fracture toughness. at least during the initial phase of the main treatment - there is no tip propagation.
As usual in the petroleum engineering literature, these models are presented in the Rather an existing fracture is reopened. Very often the treating pressure is essentially
form of proportionalities. A present collection of such proportionalities given by the same for the reopening as it was for creation of the original fracture. Since in
Hagel and Meyer [38] is reproduced in our Table 10.2. (The connection between the reopening the tip does not propagate, no tip retardation effect can be manifested.
last entry of the table and our Example 10.1 is obvious.) A possible resolution of the reopening paradox was provided by Shlyapobersky
Field tests showed that apparent fracture toughness is an order of magnitude and Chudnovsky [31). They suggest that when a hydraulically induced fracture has
larger than the laboratory value. Moreover, it is rate dependent and increases with closed and is "reopened", the same main hydraulic channel is unlikely to be formed.
the fracture length [31). Variations of fracture toughness are usually treated within A new main channel is formed during each injection, and a new process zone (damage
the R-curve approach (Bazant et al. [39]) in fracture mechanics, but extrapolation of zone) is created.
an R-curve to hydraulic fracturing scales seems to be impossible. Since no physically
sound scale law is available to predict the variation of the fracture toughness with
fracture dimension, the "calibration" experiment should be as close to the actual 10.4 Continuum Damage Mechanics in
treatment as possible. Hydraulic Fracturing

10.3.4 .Process Zone Concept Perhaps the conceptually most difficult issue in modeling fracture propagation lies
in understanding the problem of different scales [10]. In processes such as hydraulic
In a review paper by Shlyapobersky and Chudnovsky [31}, a promising train of fracturing, at least two different scales have to be considered simultaneously, namely
thought is outlined concerning the tip effects present in hydraulic fracturing. Two the scale of the fracture (tens or hundreds of meters) and the scale of the inhomo-
geneities or microcracks (one meter or less). Continuum damage mechanics (CDM)
Table 10.2 "Toughness-dominated relations" from Hagel and Meyer [38J is a convenient vehicle to handle the communication between the scale levels.
Model XI (X w(x p. (X

10.4.1 Tip Propagation Velocity from CDM


PKN [K/chj2
'V ] [K/~t] [KIC]
hIt
GDK [~r3 [Ktc
hvf r3 [K4 h
._..K_L f3 The engineering (or macro-) scale in hydraulic fracturing is of the order of one
hundred meters. With respect to this scale the damage due to microcracks can be
KICh
r
f 12 E'V considered continuous. The material damage at a given point is characterized by

Sneddon (Radial) ['Vf S


[K7c V] 1/5
[K~c s
the dimensionless variable D. It varies between zero (undamaged state) and unity
(total damage). The evolution of damage is described by a constitutive equation,
KIC '4 E'V
the Kachanov law [12} which relates the damage growth, dDjdt, to the net section
248 Fracture propagation Continuum damage mechanics 249

stress, cr,,, in a power law form: pronounced with larger ratio x f /1, and (2) the presence of damage does not affect
the stress field far from the damaged area.
dD Assuming linear damage growth (K = 1), substitution of the specific stress distri-
-=C~ (10.23)
dt n' bution into Eq. 10.25 and applying the rupture criterion yields the following form of
the tip propagation velocity [43]. '
where the Kachanov parameter, C, and the exponent, k, are material properties,
at least with respect to the considered scale. According to Rabotnov's interpreta-
tion [41], if a material point is partly damaged, only (1 - D) fraction of an elementary Uj----
-2 (
CI Kl_n
---
)2 (10.27)
section is able to carry the load, and hence the net section stress is defined as - rrcrH.rnin I + xf
where the product of the Kachanov parameter and the square of the scale parameter,
(10.24) -?
Cl", will playa central role.

where a is the nominal stress calculated without considering damage. Substituting


Eq. 10.24 into Eq. 10.23, the damage accumulation equation is given by 10.4.2 CDM-NK Model

dD
--C (cr)k
-- (10.25) A simple estimation of the nominal stress intensity factor is Pn.[xjZ, where Pn,f is
dt - I-D the net pressure at the fracture tip. If this approximation is substituted into Eq. to.27,
the resulting condition
Equation 10.25 was used by Wnuk and Kriz [42) in a local manner to describe
macrocrack propagation in a damaged material. It is reasonable to assume that in
the intact state of the rock D = O. As the macrocrack approaches a given material Uj = .si: (_xyz
I +x
rrcrH,min
)2 f
W;=Xj uo.zs)
point, the actual stress increases because the macrocrack acts as a stress concentrator.
In the mean time, the local damage is increasing according to Kachanov's law. The can be used as a boundary condition replacing the "net pressure equals zero" boundary
material point joins the macrocrack when the damage reaches its critical value, D = 1. condition in the two-dimensional differential model given by Eqs. 9.67 to 9.69.
Assume for the moment that the stress distribution moving ahead of the propagating Remarkably, the propagation velocity explicitly depends on the minimum principal
crack is known. This determines the velocity of the tip because a tip velocity that stress and is inversely proportional to the fracture length, at least for long fractures.
is too fast would not leave enough time for the damage to reach its critical value (In Chapter 11 we will discuss the "stress intensity paradox" connected with the
while a tip velocity that is too slow would cause the damage to evolve beyond its propagation of elongated elliptical three-dimensional cracks. From those results it
critical value without the rupture of the materiaL In Valko and Economides [43,44) will be clear that the scale parameter, i, has another easily acceptable interpretation
the stress ahead of the moving fracture was assumed to be of the form as well. For a homogeneous and perfectly elastic formation it can be substituted by
the height of a well contained elliptic fracture. In that case the stress intensity factor
(10.26) at the tip calculated from an analytical solution will be indeed proportional to the
inverse of the square root of the fracture length.)
It is convenient to consider the model in dimensionless form (see Eqs. 9.75 to
where KI.n is the nominal stress intensity factor calculated from horizontal plane 9.79). The dimensionless form of Eq. 10.28 is given by
strain approximation and I is a scale variable measured in meters. It represents the
scale of damage and can be considered as the "average distance of microcracks". -2
1/2
XfD
) 2
2
The factor l/ (I + x f) in Eq. 10.26 can be interpreted as the reciprocal number of UfD = CDiD WD.x=Xj (10.29)
( ID+xfD
the microcracks crossed by an elongated fracture. If the fracture is short with respect
to I, the stress-ahead function is the one known from elasticity theory (see Chapter 2). where the relation between the Kachanov parameter, C, and its dimensionless coun-
If the fracture is long, the stress is much less than it would be if calculated from terpart, CD. is given as
the nominal stress intensity factor because the presence of microcracks decreases _ crH,minC
C- _2 D, (10.30)
the stress concentration capability of the macrocrack. Equation 10.26 expresses a Cz'4
coupling between damage and stress for massive hydraulic fracturing in a highly
simplified form. It should be considered as a working postulate reflecting the two
and a similar relation for I is simply
main known features of the stress-ahead function: (1) the effect of damage is more I = clID (10.31)
250 Fracture propagation Continuum damage mechanics 251

The constants Ch C2, and C4 are the Nordgren-Kemp constants introduced in 10


Section 95.1. C /2
The CDM version of the Nordgren-Kemp model (CDM-NK) can be solved by DD
a finite differences shooting algorithm [44]. The effect of the two CDM parameters .'-
10-' 0.0001
can be summarized as follows:
1. For a large fracture (small damage scale), when TDxfD the dimensionless ~
solution path is affected only by the combined parameter CDT~.Figures 10.2 to 10.5 0.001
show computational results obtained for this case. 10-2
0.1 t 0.01
-2
(a) If > 1, the fracture propagation is not restricted by the rupture process
Col D
and the solution path does not differ from the original Nordgren-Kemp, In
I 111111111! 11111I
10-3 101 10S
1()3
to
1Q2
Figure 10.4 Fluid efficiency vs. dimensionless time, depending on the combined parameter
10' (After [43])

X'D
100 ~c------------------------------------
18
10-'
16
10-2 a:~ 14
1Q-3 ~ 12
10
1Q-4 .:
10-3 10-1 101 1()3 10S 107 ~ 8

to
&6
4
Figut'e'10.2 Solution of the CDM-NK model, Dimensionless length vs. time, depending on the 2
combined parameter (After [43])
o ~~--~--~--~~~----~~~--~--~~
0.0001 0.001 0.01 0.1

0.0001
C j2 Figut'e 10.5 Relative net pressure increase with respect to unretarded propagation. Dependence
D D
V 0.001 on the dimensionless combined parameter and dimensionless time.

.>: cf2
DD other words, the CDM model automatically reduces to its ancestor model if the
V 0.01 ~ combined parameter is large enough.
100 L 0,'
..__1 -- (b) If CDI; < 1, the fracture propagation is retarded by the rupture process and
hence the length will be less and the wellbore width and efficiency (ratio of
V the fracture volume to the fluid volume injected) will be greater than the corre-
10-1 I 11111 sponding Nordgreri-Kemp value. As a consequence, the net treating pressure
10-3 10-1 101 10S 107 will be a multiple of the Nordgren- Kemp value. The factor by which the net
to pressure is higher than the unretarded value depends on the dimensionless time
as well. Figure 10.5 shows the relative pressure increase with respect to the
Figure 10.3 Dimensionless maximum width (or net pressure) vs. time, depending on the combined
parameter (After [43J) unretarded value.
252 Fracture propagation Continuum damage mechanics 253

where
2. When TD cannot be neglected with respect to XfD, the corresponding CDT;
pressure-time solution path is reached asymptotically for late times, while in early ko = min(CDI~, 1)
times the curve has a sharp decreasing character [44].
kl = 0.62 + 0.38 x k5
As seen from Figure 10.5, the combined parameter CDT~ can be used to describe
k - k-O.37
the. "abnormally high pressures" during the whole fracturing treatment. For large
scale treatments it is reasonable to assume that the fracture is large with respect to
2 - (10.36)

the damage scale, and hence, the only relevant parameter is the combined param- E'h4 CB) 1/3
-2
eter, CDlD'
k3 = 4.75 x ko x
( --J-I=.
IJJ1-
If the combined parameter is relatively high, tip propagation rate is controlled by
The logic behind the formulas is simple. The factor IcDM reduces to unity if ko
the viscosity of the fluid (similar classical fracturing models.) When the combined
is unity, i.e. the tip propagation is unretarded. For ko = 0, the factor YCDM reduces
parameter is low, propagation is retarded by the damage evolution, and hence the
to the PKN value, i.e. 7r/5. Thus, the CDM-PKN model reduces to the PKN model
fluid viscosity has very limited effect on the propagation rate.
exactly if ko is set to unity. Once tip retardation is significant, i.e, the value of ko
becomes less than unity, the factor ICDM starts to grow. However, for very large
10.4.3 CDM-PKN Design ModeJ dimensionless limes (long fractures) the ICDM factor reduces to unity because all
the curves on Figure 10.2 to 10.4 join the unretarded curve at very late times. The
While the CDM theory resulted in the introduction of two new parameters, namely variable k3 in the correlation is designed to ensure this "late-time" convergence.
the Kachanov parameter, C, and the average distance between microcracks, T, a In the typical design mode, the desired fracture length x f is specified. The PKN
~arge-sca~~treatment is affected only by the combination of these two parameters, width, Ww.O.PKN, may then be calculated from Eq. 9.14 and, considering CDM,
I.e. by C I . From now on we restrict our consideration to such large treatments and adjusted according to Eq. 10.32. The average width can be obtained from Eq. 10.33.
use only the combined parameter to characterize the retardation effect. The subsequent steps follow a conventional design procedure presented e.g. by
Careful examination of the results of the CDM -NK model shows that a small value Meng [47].
of the combined parameter, C DI~, has two effects on the net pressure: It elevates The CDM-PKN model is especially useful for the analysis and design of massive
the level of the net pressure curve and it causes a more uniform distribution of net hydraulic fracturing with low viscosity fluid (e.g. water) where the observed high
pressure inside the fracture (Economides and Valko [45]). net pressures cannot be related to viscous dissipation.
It is possible to account for these two effects directly in a simplified design
version. Seiler [46] fitted the results of the CDM-NK model and obtained a simple
Example 10.2 CDM-PKN Simulation of a Water Fracturing
modification of the PKN model. The CDM version of the PKN model (CDM-PKN)
is defined by the two equations Treatment
Determine the created fracture length, maximum wellbore width, average width and
Ww,O = I CDMWw.O.PKN (10.32)
fluid efficiency for a waterfrac treatment from the basic data given in Table 10.3. First
and apply the PKN model without tip effect and find the propagation pressure at the end
(10.33) of pumping. Then repeat the calculations using the additional information that the
propagation pressure at the end of pumping is 63_2 MPa (9180 psi).
where ww.O,PKN is the maximum wellbore width calculated from the PKN width
equation (Eq. 9AO). The factor I CDM is introduced to account for the larger width Table 10_3 Formation and fluid characteristics for
at the wellbore and the factor YCDM is somewhat different from the one used in the Example 10.2
PKN model (i.e. in Eq. 9A1) because the distribution of the pressure, and hence, the ' 8.5 x 1010 Pa 1.2 x 107 psi
variation of the width, is different from the PKN case. Both these factors depend on hi 40 m 131ft
the combined parameter, C D1b. The Seiler correlation is given by P. 55 MPa 7980 psi
CL 8.5 x 10-5 m/sl/2 0.0022 ft/min 1/2
(1034) Sp 0.015 m 0_59 in
JL 0.001 Pa s lcp
and 0.016 m3/s 24_2/2 = 12.1 bpm

YCDM =
7r4' + (7r"5 - 4'7r) ko, (10.35) t, 6000 s 100 min
...

254 Fracture propagation Continuum damage mechanics 255


Table 10.4 PKN-a simulation without tip effect for Example 10.2 Table 10.5 CDM-PKN simulation for Example 10.2
XI =:48.7 m(160 ft) T/ 1.3%
=:
XI = 43.9 m(144 ft) T/ = 11.1%
w ..o = 0.0010 m(O.04O in) P"'pr = 56.1 MPa(8130 psi) pw.pr = 63.2 MPa(9170 psi)
w..o =: 0.0078 m(O.31 in)
w =: 8.48 x 10-3 m(O.00064 in W = 0.0061 m(O.24 in)

Solution 6000

Calculations similar to Example 9,4 give the PKN-a results which are shown in
Table 10,4.
Clearly, the PKN propagation pressure is very low compared to the value given as
xf = -'----{7'(
4"[1 + 7 x e
~-----:-:-:--:-1/4} .
0.016 x

-0.00039I/
]
40
(0.016 x 0.001 x Xf)
8.5 x 1010

known in this example. This happens often when the fluid is water because most of the + 2 x 8.5 x 1O-5J6000 x go 0) + 2 x 0.015
hydraulic fracturing models consider only viscous dissipation of energy, and water has
a very low viscosity. The likely explanation of the high propagation pressure observed where gO~ =: 1.415. The solution of the equation and the other results stemming from
in practice is that the propagation of the tip is retarded. Within the CDM-PKN model, it are presented in Table 10.5.
this would mean a smaller than unity value of the combined parameter CDlb. A trial Compared to the results in Table 10,4, the length is smaller and the width and net
and error procedure can be used to reveal the value of the dimensionless combined pressure are considerably larger than the PKN values. The retarded propagation allows
parameter, which gives the specified propagation pressure, 63.2 MPa (9180 psi). Here less surface for leakoff and spurt loss, so the fluid efficiency increases compared to the
only the last iteration is shown. Assume that the current estimate of the CDlb parameter unretarded case.
is 0.001. According to Eq. 10.36 the constants leo, kb k2 and k3 are obtained as It is useful to calculate the combined parameter, ct',
characteristic for the formation,
from its dimensionless value, 0.001. Combining Eq. 10.30 and 10.31 we obtain
leo =: min(CDlb. 1) = 0.001
kl = 0.62 + 0.38 x kJ = 0.62
k2 = ko O.37 = 12.9 Substituting the closure pressure into the minimum principal stress and using the expres-
sions for coefficients Cj, Cz and C4 given by Eq. 9.76, the following value is obtained:
k3 = 4.75 x ko x (E'~}Cf)
u: 15
= 4.75 x 0.001
-2 ( ih f ) 2/3 -2
Cl = 0.0992 x jJ.CLE'z x Pc x CDID
x 8.5 x 1010 x 404 x (8.5 x 10-5)8] 1/3
= 0.00039
[ 0.0165 x 0.001 0.016 x 40 )2/3 7
= 0.0992 x ( 0.001 x 8.5 x 10-5 X (8.5 X 1010)2 x 5.5 x 10 x 0.001
From Eq. 10.34, = 5.5 x 10- m j(pa. s) [0.042 ftZj(psi s)].
7 2

The dimensioned combined parameter clz =: 5.5 x 10-7 mZj(pa s) [0.042 ftz/(psi s))
is considered to be a property of the formation. 0
where Xt is still unknown. From Eq, 10.35, y = 7'(/4. The real power of the CDM-PKN theory is the ability to extrapolate to other
From Eqs. 9,40, 10.32 and 10.33, conditions. In the following example we calculate the propagation pressure at the
end of another 100 minute treatment with injection rate twice the original value.

Example 10.3 CDM-PKN Simulation with Increased Injection Rate


1/4
_ -O.00039xf (0.001 x 0.016 x xf
- [1 + 7 x e ]x 8.5 X 10iO )
Determine the created fracture length, maximum wellbore width, average width, fluid
efficiency and propagation pressure for a waterfrac treatment if the basic data is the same
as in Table 10.3 of Example 10.2 except with a double injection rate, i =: 0.032 m3js
Equations 9,40, 9,46, 10.32 and 10.33 can be reduced to one equation, eliminating all per one wing (48,4 bpm per two wings). Use the combined CDM parameter obtained
variables but xf' Thus, in Example 10.2, i.e. cr =
5.5 x 10-7 m1/(Pa. s) [0.042 ftz/(psi s)).
256 Fracture propagation Pressure decline analysis and tip retardation 257

Table 10.6 CDM-PKN simulationwith increased injection 2E'V; 4E' ( 4' )


rate,Example 10.3 Pw=Pc+-h, --h Sp- hCL~ xg(LltD,a)
T! fX f T! f T! f
Xj = 8.1 m(268 ft) = 17.2%
w w.o = 0.013 m (0.51 in)
I)

P ....p, =
68.7 MPa (9980 psi) = b + m x g(LltD, a). (10.37)
W = 0.010 m (0040 in)
From Eq. 10.37, it is seen that the leakoff coefficient can still be obtained from the
slope, m, of the straight line on the g-plot as
Solution T!hf
CL = --(-m). (10.38)
4...;t;E'
When the injection rate is 0.032 m3/s, the dimensionless parameter, CDI~, will be
different from its previous value, 0.001. It is easily seen that for an injection rate An additional relation between the fracture half-length and spurt loss coefficient
twice the previous value the dimensionless combined parameter will be less by a exists in terms of the intercept, b, of the straight line as follows:
factor of 2-2/3. Thus Col; = 0.001 x 2-2/3 = 0.00063. With this new dimensionless
combinedparameter,the solutionof the CDM-PKNmodelgives the results summarized s _~ 7Th/(pc - b)
(10.39)
in Table 10.6. P - lxfhf + 4E'
Compared to the results in Table 10.5, the increased injection rate brings about an
increased propagationpressure and increased efficiency.0 Equations 10.38 and 10.39 are not affected by the fact that a shorter length is created
during the injection period.
It is easy to see that the second term on the right-hand side of Eq. 10.39 is
10.5 Pressure Decline Analysis and Tip Retardation negative. Since the spurt loss cannot be negative (a flow into the fracture during
fracture propagation is absolutely not possible), the fracture half length, x f' should
While the continuum damage mechanics model may not be the ultimate solution to be less than a maximum value, xf,max, where
the challenging problem of modeling hydraulic fracture propagation, it provides a
framework to predict fracture performance under variable conditions while avoiding 2E'Vi
(10.40)
some unreasonable features (such as independence or decrease of treating pressure Xf.max = 7Th}Cb- Pc)'
with increasing injection rate) often inherent in other theories. However, some ques-
tions (relevant to any practical theory of hydraulic fracturing) have to be answered. On the other hand, the propagation pressure provides an estimate of the fracture
length if the propagation is unretarded. Rearranging Eq. 9.89 we obtain the unre-
1. Can the tip retardation phenomenon be revealed from field data? tarded fracture length, xf,ur, as
2. Can the combined parameter be determined from an injection test?
3. If the combined parameter is known, can we predict the main characteristics of 4 ( hite )
Xf.ur = 0.140(Pw.pr - Pc) '3J1.V; . (10.41)
the fracture created in the main treatment? Can we do this even if some of the
parameters (e.g. injection rate) are different in the calibration test than in the
main treatment? If the unretarded fracture length estimate is longer than the maximum possible value
given by Eq. 10.40, then the original PKN width equation (i.e. Eq. 10.41) did not
Some of these questions were answered by the previous examples. We saw how hold during propagation and the tip was retarded. The inequality
the tip retardation manifests itself as high propagation pressure once we know the
parameters of the fluid loss process. We saw how to determine the combined param- Xf,ur > xf.max (10.42)
eter, and how to apply it in design (even if some of the treatment characteristics are
different from the calibration test.) is a clear indication of tip retardation.
What is still not clear is how to obtain the leakoff, spurt loss and tip retardation Note that the viscosity of the fluid is in the denominator of Eq. 10.41, so for thin
characteristics simultaneously. The remaining part of this chapter is devoted to this fluids, the basic inequality 10.42 will often signal tip retardation.
principal issue. Once the phenomenon of tip retardation is established, the use of a model capable
The essential idea is that the presence of tip retardation does not change the laws to reproduce the phenomenon is a necessity. One of the models providing the descrip-
of linear elasticity and leakoff which are in effect during the decline period. For tion of tip retardation is the CDM-PKN model which uses the combined parameter,
the PKN geometry, i.e. for the vertical plane strain assumption, Eq. 9.105 is still c12 This parameter modifies Eq. 10.41 and keeps the length below the limit posed
in effect: by Eg. 10.40.
258 Fracture propagation Pressure decline analysis and tip retardation 259

10.5.1 Resolving Contradictions with Continuum 3. The fracture volume at the end of injection equals the volume after the after-
Damage Mechanics growth process plus the spurt loss during the after-growth process:

In the following we suggest a procedure to obtain a set for parameters of the CDM- WISIXj + 2Sp~xj = WprXf.pr'
PKN model which reproduce the main characteristics of an injection test without
inner contradiction. The procedure is counterpart to the pressure decline analysis 4. Eqs. 9.40 and 10.32 give the well bore maximum width and Eq. 10.33 the average
suggested in Section 9.6.2. width at the last instant of propagation.
We start from the straight-line fit of the g-plot. Once m and b are determined, 5. Net wellbore pressure calculated from Eq. 9.43 (with the wellbore width obtained
the leakoff coefficient, C L, is obtained from Eq. 10.38. The next step is to calculate from Eq. 10.32) should equal the observed propagation pressure minus the
Xj,max from Eq. 10.40 and Xj,ur from Eq. 10.41. If the basic inequality 10.42 does closure pressure.
not hold, there is no reason to use a tip retardation model and the procedure of
Section 9.6.2 should be followed. If the basic inequality holds, the fracture propa- The algorithm to determine the parameters is numerically simple:
gation is interpreted in terms of the CDM-PKN model. The following parameters
1. Assume a dimensionless combined parameter, CDlb, less than unity.
are unknown:
2. Calculate
-2
1. the combined parameter, Cl ,
ko = CDl1
2. the spurt loss parameter, S p and
3. the after-growth, ax], kl = 0.62 + 0,38 x 14
In addition, we will use the following variables: k2 = kO O.37

4. the half-length at the last instant of propagation, x j,pro E'h4 C8~) 1/3
k3 = 4.75 x ko x fs L e
5. the average width at the last instant of propagation, Wpro and ( VifJ-
6. the average width following the short after-growth process, WlS!.
_ (b+ 1.415 x m - pc)rrhj
WISI = ------___;;'---~
It is assumed that the fracture propagates according to the CDM-PKN model 2E'
(Eqs. 10.32 to 10.36) combined with the a-type material balance (Eq. 9.46). In _ 2rrhf(0.25 - 0.05ko)(Ppr - Pc)
addition, it is assumed that the fracture propagates after the pumps stop, with the wpr ==
E'
incremental length given as the after-growth parameter. It is assumed that the after-
growth happens in a short time; this assumption has to be checked after the model 3. Solve the following nonlinear equation for Xj.pr
is fitted. The last assumption means that the pressure after shut-in is modeled by
Xj.pr
Eq. 10.37, where Xf is the sum of the half length at the last instant of propagation
plus the after-growth.
Input data required for the estimation procedure are the following:
_ { wpr }4 (Elte)
- 3.27 x [1+(k1k2-1)exp(-k3Xj,pr)] x [rr(0.25-0.05ko)] fJ-Vj'

1. formation data: E', hj, and Pc; 4. Solve for ~x j the following quadratic equation (select the positive root)
2. treatment data: Vj, i, and u;
3. observed data: m and b of the g-plot straight line and the propagation pressure, _ _ _ 2rrhj(pc - b)~xj ~Xj Vi
WISrxf.pr - WprXj.pr + wISJ~xf + ,+ = O.
Ppr' 4E (xf.pr+~j)hj
The following conditions have to be met:
5. Calculate the spurt loss coefficient from
1. WISI satisfies
_ (b + 1.415 x m - Pc)1r:hj
wISI = 2' .
2. x j and ~X j are related through A parameter set is accepted if both and ~X f and Spare positive real numbers.
(Otherwise the assumed tip retardation is still not enough to keep the propagation
length below its physically possible maximum.) It is suggested to start the procedure
--"~---.--------. --_._-_.:;;

260 Fracture propagation Pressure decline analysis and tip retardation 261

with a near unity dimensionless combined parameter, CD1b' If the parameter is Table 10.7 Consistent data sets determined from the pressure decline data
for Example 10.4
decreased step by step, the first feasible parameter set appears when the spurt loss
-,
coefficient is zero. Further reduction of the combined parameter will yield additional CDr; Sp (m) x{.pr (m) 6.xf (m) ~~Sl (m) ~pr (m)
feasible parameter sets. This is not surprising because we posed five constraints on
0.12 0.000356 211 116 0.0044 0.00720
six variables. The following example illustrates the procedure and addresses the issue
0.11 0.00126 187 76.1 0.0044 0.00722
of uniqueness. 0.10 0.00231 163 51.3 0.0044 0.00723
0.09 0.00359 140 34,4 0.0044 0.00725
0.08 0.00526 118 22.7 0.0044 0.00726
Example 10.4 Pressure Decline Analysis with the CDM-PKN Model 0.D7 0.00750 97.1 14.4 0.0044 0.00728
0.06 0.0107 77.5 8.72 0.0044 0.00729
Consider the formation, fluid, injection and pressure decline data analyzed in 0.05 0.0155 59.3 4.88 0.0044 0.00731
Examples 9.7 and 9.8. Repeat the pressure decline analysis without artificially changing 0.04 0.0234 42.7 2.44 0.0044 0.00732
the viscosity, i.e, use the originally accepted estimate, f.J =
0.1 Pa s (100 cp). Account 0.D3 0.0384 27.9 1.01 0.0044 0.00734
for possible tip retardation. 0.02 0.0738 15.3 0.298 0.0044 0.00735
om 0.215 5.5 0.0377 0.0044 0.00737

Solution
that the total after-growth time is 6.ta.o =
10 min, where we use an extra subscript,
The slope and intercept of the straight line are m = -1.208 MPa and b = 44.32 MPa. "0", to denote that this value is observed. Using the CDM-PKN model, we make a
The leakoff coefficient is the same as in Example 9.7, i.e, CL = 1.70 X 10-5 m/s1/2 simulation run for every parameter set (Col;, CL, Sp) shown in Table 10.7. At the end
(4.32 x 10-4 ft/mini(2). of pumping we calculate the fracture propagation rate from the length increment during
The maximum possible length is determined from Eq. 10.40. the last minute of injection, and denote it by U f. Assuming that this propagation rate will
linearly decrease to zero, the calculated after-growth time is given by Ata 2Axt/uf. =
_ 2E'Vi _ 2 x 6.13 X WiD x 125 _ 362 (1190 f) We accept the data set for which the calculated and observed after-growth time differs
xfmax- , - 6- m t I.
. TChj(b - Pc) tt X 5.182 x (44.32 - 39.3) x 10 the least, i.e. we minimize the deviation of 6.ca from Ala.a.
The unretarded propagation length from Eq, 10.41 is In the given example, the procedure leads to the data set corresponding to CDl~ =
0.07, Sp = 0.0075 m (0.30 in) and 6.ta = 11 min. The suggested criterion to select the

xfur = 0.140(Pw.pr - Pc)


4 ( hite )
E'3f.JVi
appropriate parameter set is rather selective. If the previous parameter set (Cot; 0.08) =
is selected, the after-growth time is too large (Ata =
14 min) and if the next parameter
51.84 x 2220 set (CDI~ = 0.06) is selected, the after-growth time is too low (Ata 8.5 min).=
= 0.140 x (5.56 x 106)4 0 3 = 742 m(2440 ft). The combined parameter in dimensioned form is obtained from
. (6.13 X 101 ) x 0.1 x 125

Since xf.ur > xf.max, the tip propagation is retarded and the CDM-PKN model should c72 = 0.0992 x ( Vih f ) 2!3 -2
f.JCLE'2te X Pc X Colo
be used.
Following the procedure delineated above, we start the calculation with a dimension-
125 x 51.8 ) 2/3
less combined parameter equal to unity and decrease it step by step. The first positive
-2
= 0.0992 x (
0.1 x 1.7 x 10-5 X (6.1 X 1Ol0)2 x 2220
x 39.3 X 106 x 0.07
real root of the quadratic equation for 6.xf appears when Colo =
0.12. The results are
shown in Table 10.7. = 1.62 x 10-5 m2J(Pa s)[1.2 ft2/(psi s)].
The obtained parameter sets describe the pressure behavior equally well during the
injection period. In the time period after the pumps stop the parameter sets still result in The results of the parameter estimation procedure are summarized in Table 10.8.
an equally good match of the pressure decline curve when only that part is considered
which gives a straight line on the g-plot. Table 10.8 Accepted parameter set for
Additional considerations are needed to select one of the above parameter sets. One Example lOA
possible procedure is based on the early behavior of the pressure decline curve. The
CL = 1.7 x 10-5 rn/Slf2(4.3 x 10-" ft/minl/2)
technique is the following. The approximate time of after-growth is estimated from the
Sp = 0.0075 m(0.30 in)
g-plot. From Figure 9.2 we see that for the SFE3 injection test approximately 10 minutes
C/2 = 1.6 x 10-5 m2/(Pa 5) [1.2 ft2/(psi. 5)]
can be considered as a potential time interval for after-growth because the observed
pressure points are located markedly off the straight line in the first 10 min. Assume
=
AtQ 11 min
262 Fracture propagation References 263

Table 10.9 CDM-PKN simulation result at the end of the injec- propagation pressure at the end of pumping and the decline pressure curve are matched
tion for Example lOA correctly. At early times the treatment pressure is overestimated. Reasons for this might
= 98.3 m(323 ft) be numerous and are beyond the scope of this chapter. Any attempt to "improve" the fit
Xf.pT TJ = 29.7%
w,.,.o = 0.00943 meO.37! in) uf = 0.039 m/s(7.S ft/min)
by reproducing the small but evident changes in the trend of the pressure decline curve
iii = 0.0073 m(0.287 in) pw.pr = 44.88 MPa(6509 psi) would be meaningless because these capricious changes are due to systematic errors of
.the particular measurement system. Also, we recall that the purpose of this exercise is
to illustrate the flexibility of the tip retardation concept. The results depend heavily on
Table 10.10 Fracture length variation the selected values of plane strain modulus, closure pressure and height. The decision
during the after-growth to accept, discard or modify these values is an important engineering activity, but is
process in Example 10.4 also beyond the scope of this chapter.
I, min Xj
=
At this point it is in order to recall that the selected viscosity, I-' 0.1 Pa s (100 cp),
corresponds to average shear rate conditions in the fracture with average width of 8 mm
37 9S.3 m 323 ft (0.3 in) as seen from Table 9.15. Since the final width shown in Table 10.9 is 7.3 mm
38 101 m 330 it (0.29 in), there is no reason to adjust the viscosity further.
40 105 m 344 ft The simulation stops when the pressure decreases below the closure pressure (i.e. the
42 lOS m 354 ft width reaches zero.) The remaining part of the pressure decline curve (after 180 min)
44 110 m 362 ft cannot be simulated with the Carter leakoff concept. If we wish to describe that period
46 111m 364 ft as well, the detailed leakoff model of Section 9.5 should be used. It is obvious that the
from 48 111 m 365 ft
CDM-PKN concept can be applied together with any fluid-loss model, including the
particular one of Section 9.5. 0
Using the treatment data and the parameters in Table 10.8 we simulate the injection
test as follows. During the injection period, the CDM-PKN width equation (Eqs. 10.32,
involving Eqs. 10.33, 10.34, 10.35, 10.36 and 9.40) is solved together with the material References
balance Eq. 9.46. The results, shown in Table 10.9, are obtained for the end of the
injection period. 1. Nolte, K.G. and Smith, M.B.: Interpretation of Fracturing Pressures, J. Petrol. Technol.,
Taking into account that the after-growth process lasts about 11 min and the propaga- 1767-1775,1981.
tion rate decreases linearly during this period, the fracture lengths shown in Table 10.10 2. Von Kinnan, Th.: Festigkeitsversuche unter allseitegem Druck, Z. Ver. Deutscher Inge-
are easily determined from the formula of constant deceleration. nieure, 55 (42),1749-1757,1911.
Now we have. all the information needed to use the first line of Eq. 10.37 with 3. Timoshenko, S.P.: History of Strengths of Materials, McGraw Hill, New York, 1953.
the partly varying lengths given in Table 10.10 and the pressure decline can be easily 4. Gramberg'J.: A Non-conventional View on Rock Mechanics and Fracture Mechanics,
simulated. The simulated total pressure curve is shown in Figure 10.6. Note that the Balkema, Rotterdam, 1989_
5. Coulomb, C.A: Sur Une Application des Regles de Maximis et Minimis a Quelques
4S 120 Problemes de Statique Relatifs a I' Architecture, Acad Roy. des Sciences Memoires de
X E math. et de physique par divers savans, 7, 343-82. 1773.
!IS
Q.
44 100 X 6. Mohr 0_: Welche Umstande bedingen die Elastizitatgrence und den Bruch eines Mate-
~ rials? Z. VeT. dt. Ing., 44, 1524-1530, 1572-1577,1900.
3:: C.
t::
0. 43" 80
~ 7. Griffith A.A. The phenomena of Rupture and Flow in Solids, Phil. Trans. Royal Soc.
e e
.....
:::l
42 60
::::I
U 8.
London, Ser. A 221, 163-198, 1920.
Griffith AA The Theory of Rupture Proc. First International Congress on Applied
l!!
c. : Mechanics, Delft, 55-63, 1924.
e <11 40 "'C
9. Jaeger J.C and Cook N.G.W.: Fundamentals of Rock Mechanics, Chapman and Hall,
s ~
"5
0
London, 1976.
~ 40 20 iii 10. Lamaitre J.: Local Approach of Fracture, Engineering Fracture Mechanics, 25(5/6),
pc, dosure pressure o 523-537, 1986.
39 '0 11. Boone T.J., Wawryznek, P.A and lngraffea, AR.: Simulation of the Fracture Process
0 50 1 150 200 Zone in Rock with Application to Hydrofracturing, Int. J. Rock Mech. Min. & Geomech.
Time, min Abstr., 23(3), 255-265, 1986.
12. Kachanov L.M.: Time of Rupture Process under Creep Conditions, Izv. Akad: Nauk SSR,
Figure 10_6 Observed and calculated pressure curve for Example 10.4 Otd. Tekh., 8, 1958.
,.

References 265
264 Fracturepropagation

33. Cleary, M.P. Wright, C.A Wright, T.B.: Experimental and Modeling Evidence for
D. Chudnovsky A: Crack Layer Theory, NASA CR-17463 , Case Western University. Major Changes in Hydraulic Fracturing Design and Field Procedures, Paper SPE 21494
Cleveland,OH, 1984.
presented at the Gas Technology Syrnp., Houston, TX, Jan. 23-25, 19~1. .
14. Mott N.F. Engineering, 165, 16-18, 1948_ Van den Hoek, PJ., Van den Berg, J.T.M., Shlyapobersky, J.: Theoretical and Experi-
34.
15. Billington, E. W. and Tate, A: The Physics of Deformation and Flow, McGraw Hill, mental Investigation of Rock Dilatancy Near Tip of Propagating Fracture, Int. J. Rock
New York, 1981. Meeh. Min. & Geomech. Abstr., 30, 1261-1264, 1993.
16. Zienkiewicz, O.C.: The Finite Element Method in Structural and Continuum Mechanics, Gardner, D.C.: High Fracturing Pressures for Shales and Which Tip E~e~~s May Be
35.
McGraw Hill, London, 1976. Responsible, SPE paper 24852 presented at Techn. Conference and Exhibition, Wash-
17. Erdogan, F. and Sih, G.C: On the Crack Extension in Plate under Plane Loading and ington, D.C., Oct 4-7, 1992. . .
Tranverse Shear, J. of Basic Eng., ASME, 85, 519-527, 1963. Shlyapobersky, J. Wong, G.K and Walhaugh, W.W.: Overpressure-Calibrated Design
36.
18. Sih, G.C: Strain-Energy-Density Factor Applied to Mixed-Mode Crack Problems, Int. of Hydraulic Fracture Simulations, Paper SPE 18194 presented at the 63rd Technical
J. Fracture Mechanics, 10, 305-321, 1974. Conference and Exhibition, Houston, TX, Oct 205, 1988.
19. Boone T. J., Ingraffea AR. and Rogiers J.-C: Visualization of Hydraulically Driven Yew, C.H. and Liu G.: The Fracture Tip and K1C of a Hydraulically Induced Fracture,
37.
Fracture Propagation in Poroelastic Media, J. Petrol. Technol., (June), 574-580, 1969. Paper SPE 22875 presented at 66th Ann. Techn. Conf. and Exh. of SPE, Dallas, Oct.
20. Gerstle, W. H. and Xie, M.: FEM Modeling of Fictitious Crack Propagation in Concrete, 6-9,1991. .
1. Engrg. Mech., ASCE 118(2) 416-434,1992. Hagel, M.W. and Meyer, B.R.: Utilizing Mini-Frac Data to Improve Design and
38.
21. Bouteca, M.L Hydraulic Fracturing Model Based on a Three-dimensional Closed Form: Production, J. Canadian Petro Technol., 33, 26-35, 1994. .
Tests and Analysis of Fracture Geometry and Containment, SPE Production Engineering 39. Bazant, Z.P., Gettu, R. and Kazemi, M.T.: Identification of Nonlinear Fracture Prop-
(Nov.), 445-454, Trans. AIME, 285, 1988. erties From Size Effect Tests and Structural Analysis Based on Geometry Dependent
22. Clifton Rl and Abou-Sayed AS.: On the Computation of the Three Dimensional R-curves, Int. J. Rock Mech. Min. & Geomech. Abstr., 28(1), 43-51, 1991.
Geometry of Hydraulic Fractures, Paper SPE 7943, Symp. on Low-Permeability Res. 40. Chudnovsky, A. and Gorelik, M.: Statistical Fracture Mechanics, - Basic C~nce?ts a~d
Denver, 1979. Numerical Realization, Probabilities and Materials: Tasks, Models and Applications, in
23. Thiercelin, M., Jeffrey, R.G. and Ben Naceur, K: Influence of Fracture Toughness on Breysee D. (ed.) Kluwer, Boston, 1994.
the Geometry of Hydraulic Fractures, SPE Production Engineering, (Nov.), 435-442, 41. Rabotnov, Y.N.: Creep Rupture in Proc. XII Intern. Congress Appl. Mech., Stanford,
1988. Springer, Berlin, 1969. .' .
24. Heuze, F.E., Shaffer, R.J., Ingraffea, A.R. and Nilson, R.H.: Propagation of Fluid- 42. Wnuk, M.P. and Kriz, R.D.: CDM Model of Damage Accumulation In Laminated
Driven Fractures in Jointed Rock. Part 1 - Development and Validation of Methods Composites, International Journal of Fracture, 28, 121-138, 1~85. .
of Analysis, Int. J. Rock Mech. Min. & Geomech. Abstr., 27, 243-257, 1990. 43. Valko, P. and Economides, M.J.: Continuum Damage Mechanics Model of Hydraulic
25. De Pater, C.J., Weijers, L. Van den Hoek, P-I. Barr, D.T.: EXperimental Study of Non- Fracturing, J. Petrol. Technol., 198-205, 1993. .
linear Effects in Hydraulic Fracture Propagation, Paper SPE 25893 presented at the SPE 44. Valko, P. and Economides, M.J.: Propagation of hydraulically induced fractures - a
Joint Rocky Mountain Regional Meeting and Low-Permeability Symposium, Denver, continuum damage mechanics approach, Int. J. Rock Mech. Min. & Geomeeh. Abstr.,
CO, April 12-14, 1993. 31(3) 221-229, 1994. . .
26. Medlin, W.L. Fitch, lL.: Abnormal Treating Pressures in Massive Hydraulic Fracturing 45. Economides, M.J. and Valko, P.: Interpretation and modeling of hydraulic fractu~g
Treatments, J. Petrol.Technol., 633-642, 1988. phenomena with continuum damage mechanics - An application to engineering design,
27. Palmer, tD. and Veatch Jr., RW.: Abnormally High Fracturing Pressures in Step-Rate In. Siriwardane H.J. and Zaman, M.M. (ed.) Computer Methods and Advances In Geome-
Tests, SPEPE, (Aug.), 315-323; Trans. AlME, 289. chanics, 1579-1583, Balkema, Rotterdam, 1994. .
28. Cleary, M.P.: Rate and Structure Sensitivity in Hydraulic Fracturing of Fluid-Saturated 46. Seiler, R.: Development of a Fracture Design Procedure Based on Continuum Damage
Porous Formations, Proc. 20th u.s. Rock Mechanics Symosium, Austin, TX, 124-142, mechanics, Diploma Thesis, Mining University Leoben, 1993.. .,
1979. 47. Meng, H-Z.: The Optimization of Propped Fracture Treatments, In Reservo.lr Stimu-
29. Jeffrey, R.G.: The Combined Effect of Fluid Lag and Fracture Toughness on Hydraulic lation, Economides, M.J. and Nolte, KG. (ed.) Prentice Hall, Englewood Cliffs, N.J.,
Fracture Propagation, Paper SPE 18957 presented at the 1988 SPE Joint Rocky Mountain 1989.
Regional!Low Permeability Reservoirs Symposium, Denver, CO, March 6-8.
30. Khristianovitch, S.A. and Zheltov, Y.P.: Formation of Vertical Fractures by Means of
Highly Viscous Fluids, Proc. World Petroleum Congress, Rome, 2, 579, 1955.
31. Shlyapobersky, J. and Chudnovsky, A: Review of Recent Developments in Fracture
Mechanics with Petroleum Engineering Applications, Paper presented at EUROCK'94,
Delft, August 29-31, 1994.
32. Desroches, J. Lenoach, B. Papanastasiou, P. and Thiercelin, M.: On Modelling of Near
Tip Processses in Hydraulic Fractures, Int. J. Rock Mech. Min. & Geomeeh. Abstr., 30,
1127 -1134, 1993.
11
FRACTURE HEIGHT
GROWTH (3D AND P-3D
GEOMETRIES)

Geological formations consist of more or less distinct layers. As usual, only one
or few of the layers contain hydrocarbons. The well is perforated to access the
hydrocarbon-containing target layer(s). A hydraulic fracture is intended to remain
within the perforated interval, i.e. to be contained. However, simple observations of
lithological logs frequently suggest no clear lithological barriers for fracture height
containment, although reservoir height delineation logs often suggest a much clearer
limit for the productive interval. In certain cases, lithological homogeneity may
suggest penetration by a growing fracture height into adjoining porous or even non-
porous media. Indeed, there is evidence that "the fracture may grow up, out and
down the perforated interval" (Rahim and Holditch (1D.
In the models considered in Chapter 9, strong assumptions are posed on the
resulting geometry of the fracture. These assumptions are designed to leave
only one degree of freedom besides the width, and hence the description two-
dimensional (2D).
If the height is considered to be known, the additional degree of freedom (besides
the width) is the length. The different variations of the PKN and KGD models are
such constant-height models. The constant-height models are the most typical but
not the only two-dimensional models.
The radially propagating fracture model is not of constant-height type, but still only
one degree of freedom is left (the radius); thus, it also belongs to the two-dimensional
group of models.
In contrast to the two-dimensional models, three-dimensional (3D) models describe
the change of the geometry allowing for complex shape variation which can be
determined using only three space coordinates. Three-dimensional hydraulic frac-
turing models envision the formation to consist of horizontal layers of different
thickness, mechanical properties and minimum horizontal stress. This detailed infor-
mation is combined with the characteristics of fluid flow inside the fracture (including
.----_.-- ... ,----, '---_.,-,--_., .. _------ __
.. ."._-----,--- _._---_ .._....._---

268 Fracture height growth Equilibrium fracture height 269

the resulting pressure distribution) to determine the direction and velocity with
11.1 Equilibrium Fracture Height
which the fracture boundaries move. The width at a given lateral and vertical loca-
tion is calculated from a detailed description of the stress and displacement field. Griffith's theory of equilibrium allows us to calculate the length of a line crack
The numerical methods of linear elasticity theory are used extensively. Fluid flow if the inner pressure is known as a function of location. Applying this theory to
is considered two-dimensional, assuming quasi-steady-state flow between (locally) height-determination involves some technical difficulties, e.g. complexity of notation
parallel plates. involved, possible non-existence or multiplicity of solutions to the resulting non-
Often the name pseudo-three-dimensional (P-3D) model is also used in hydraulic linear system of equations. Also, some basic assumptions considered trivial in the
fracturing (Settari and Clearly [2]). Most P-3D models do not consider vertical fluid technique need particular caution. In all of our derivations we will consider only the
flow but, instead, predict height from net pressure as a function of lateral location three-layer case. Some remarks concerning more layers will be given at the end of
and with the local use of fracture mechanics criteria. Analytical solutions for simpli- the discussion.
fied cases are used (Bouteca [3]) in place of detailed numerical methods of linear The approach of Simonson et al. [41 has been used widely in the petroleum
elasticity. In some cases the difference between three- and pseudo-three-dimensional industry. The penetration of fracture into the upper and lower layers surrounding
models is not well defined. the target layer is determined by obtaining the equilibrium height for a given net
Both in three- or pseudo-three-dimensional models the aim of departure from pressure. The equilibrium height satisfies the condition that the computed stress
two-dimensional theory is to describe the process of hydraulic fracturing with better intensity factor at the vertical tip equals the critical stress intensity factor. The latter
accuracy. Since the geometry is known in more detail, it is natural to use this infor- quantity, often called fracture toughness, is considered to be a material property of
mation to refine the description of all other elements. In particular, in 3D models the rock.
the fluid flow submodel may take into account the movement of the fluid not only
in the lateral but also in the vertical direction (2D flow). In the future, it may even
.: ;
be possible to track the motion of fluid particles in a direction perpendicular to the 11.1.1 Reverse Application of the Net-pressure Concept
fracture surface by introducing a three-dimensional description of fluid flow. Once To understand the concept of equilibrium height, some abstraction is necessary. A
this Pandora's box is open, the simple concept of quasi-steady-state flow might be
line crack is considered in the vertical plane perpendicular to the lateral direction of
dropped in favor of unsteady flow equations which include inertia effects. An addi- propagation. At a given lateral location, x, the pressure along the vertical direction is
tional step is to consider the fluid front moving somewhat behind the boundaries of
considered to be either constant (if the hydrostatic pressure component is neglected)
the fracture, i.e. to incorporate the description of the fluid film formed near to the or varying linearly with depth (a result of increasing hydrostatic pressure). Simonson
tip. With a detailed geometry, it is also possible to consider stress intensity factors or et al. [4] considered the simplest case neglecting hydrostatic pressure and assuming
similar characteristics far more complicated than the simple K[, i.e. opening modes that the material properties of the upper and lower layers are identical. They assumed
far more complex than the simple "mode I opening" . These considerations are of that the minimum horizontal principal stress has the same direction in all the three
primary importance in research where the aim is to discover the governing laws layers. In addition, the value of O"H.min in the upper layer (0"2) and in the lower layer
and/or their combined effect.
(0"3) were assumed equal to each other and higher than in the target layer (0"1). Also
In applied engineering, the aims are somewhat different. Here the primary concern the critical stress intensity factor in the upper layer (K[C.2) and in the lower layer
is not to neglect one of the main factors that affect the resulting design, while (K[C.3) were considered equal. If all of these assumptions hold, the geometry of the
microscopically accurate description of the known subprocesses is de-emphasized. equilibrium crack is symmetric.
The admirable numerical arsenal in linear elasticity computations has evolved in To calculate the stress intensity factor at the tip, Simonson et al. [41 used the
civil engineering. Prior to the finite element method, a huge amount of available concept of superposition, defining a different net pressure in the target and in the
information on the geometry of the structures was simply discarded for lack of a two other layers. This was done not because the pressure varies inside the fracture
method to apply it. The possibility of using this information was a revelation and (it does not) but because the boundary condition at infinity changes with the vertical
became the engine of development. In hydraulic fracturing, unfortunately enough, the location (aH.min is different in the target and in the other layers). Note that in the
situation is different. Here we generate the geometry through mathematical models theory of elasticity these minimum principal stresses are used as boundary conditions
consisting of partly less accurate submodels. Therefore, the accuracy of a model at infinity. The use of superposition with respect to the change in the boundary
depends less on representing the geometry in detail and more on accounting for all condition at infinity is not a standard technique of mathematical physics, and from
the significant phenomena (such as energy dissipation in the rock.) a strictly mathematical point of view it might be questionable. It may be difficult
In this chapter we give an overview of existing methods to handle the issue to visualize that a discrete jump in a condition at infinity will cause a jump in the
of height growth. In doing this, we will point out some conceptual and technical stress at the surface of the crack. Nevertheless, if we accept that the net pressure,
difficulties inherent to existing theories. calculated as the difference p - O"H.min, can be used to calculate the stress intensity
1

270 Fracture height growth Equilibrium fracture height 271

factor even in this case (where 0H,min is not constant any more), we arrive at a useful Figure 11.1 is the notation taken largely from Warpinski and Smith [4]. One minor
technique of characterizing height containment. difference is that here we use the letter c for the half-length of the vertical line crack
(to be consistent with the other chapters of this book) while those authors use the letter
a. Here we use Yw for the vertical coordinate to make sure it is not confused with
11.1.2 Different Systems of Notation
the dimensionless coordinate to be introduced later. The thickness (i.e. the height)
Here we will consider the more complex situation where of the perforated (or target) interval is denoted by hp Those authors introduce the
"geometry factors" b: and b3. In fact the "geometry factor" h3 is not an independent
1. 02 and 03 may be different but still higher than OJ, variable, but rather the difference hp - bz- We have two degrees of freedom, b2
2. the critical stress intensity factor may be different in the upper and lower layers, and c.
.~
.. and The same problem with a somewhat different system of notation is depicted on
3. the density of the fluid is accounted for. Figure 11.2. This is the system used by Ahmed [6] and Economides [10]. Here the
key variables to determine are the upper and lower height growth, Sh; and /)"hd.
The analysis of Warpinski and Smith [5] has been generalized to account for some Also, a dimensionless coordinate system is introduced in the figure. The dimension-
or all of the above variations by many authors [5-9], but often the results are difficult less variable y is zero at the center of the crack and unity at the top. The relations
to compare because of the different systems of notations. Because of the technical between the different notations are given in Table 11.1. In Table 11.2, the dimen-
importance of the results, we try to cover the basic issues using two basic systems sionless variables are expressed in terms of the key variables for both systems of
of notations accepted in practice. notations.

Table 11.1 Relation between different notations for the three-layer equilibrium problem

System of After Ref. [5) After Refs. [6-10] Dimensionless


c Notation (Figure 11.1) (Figure 11.2) (Figure 11.2)
___ __Center of
f crack
Coordinate
Center of crack
Yw
0
YR
0
Y
0
c
Center of + Yd
1 perforation b _ hp
- 2
t::.hd - tlhu
2
Y.
2
Upper height
growth e- bz Sh; 1- Y.
Lower height
growth e - (hp - b1) tlhd 1 + Yd
Figure 11.1 Equilibrium height, system of notation after [5]
Top of
perforation b2
hp - tlhu + t::.hd
y.
2
Y
Bottom of -hp - t:,.hu + t::.hd
perforation -b3 = b1 - hp Yd

...--~
-.----
L>hu
t_
Total height
Perforated
2c hp
2
+ Sh; + t:,.hd 2

Yu height hp hp Y. - Yd
o Pressure at
center of
perforation Pcp Pcp Pcp
Yd
-1 Pressure vs, P = Pcp+ P= Pcp+ p=koo+k1y
vertical location
pg (b h; - yw)
2- pg (
t::.hd - 6h.
2
. )
- YR koo = Pcp
Y. + Yd
+ pg---
Y. - Yd
Zhp
kl =-pg---
yu-yd
Figure 11.2 Equilibrium height, system of notation after [6,7]
272 Fracture height growth Equilibrium fracture height 273

Table 11.2 Dimensionless variables in terms of the key variables of the different and
systems of notation

Relation to y
hp + 6hu + l!ihd K/,lx)ttom = --'---hp
Jr(Yu-Yd)
x 11.
-1
Pn(y)~-y--dy,
1+ Y
(11.4)
Yw "" cy YR"" 2 Y
b2 zs. where the net pressure varies with the dimensionless vertical coordinate y, for two
Relation to Y. Yu"" - Yu = 1 - .,-----,--:--=--:-
c + l!ihu + 6hd
hp reasons: The minimum horizontal principal stress is different in the different layers
b2 -h zs, and the hydrostatic pressure increases with depth.
Relation to Yd Yt=--P Yd = -1+ ---=--
c hp + l!ihu + 6hd The last entry in Table 11.1 gives the pressure as a function of the vertical coor-
2b2 -h 6hd -l!ihu dinate. The net pressure will be almost the same function, except the constant term
Relation to koo koo = Pcp + pg--2-P koo = Pcp + pg 2
ko will be the difference between koo and the minimum principal stress in the given
c 2hp layer. In general, the net pressure will be a piecewise linear function (with jumps
Relation to kl kl = -pg- k1=-pg--
2 Yu - Yd at the layer boundaries). Hence, it is advantageous to determine two indeterminate
integrals (primitive functions) for the general case where the net pressure is a linear
11.1.3 Basic Equations function of the vertical location. The following two functions are defined:

The two unknowns can be determined from two equations. The two equations are (11.5)
derived from the conditions of equilibrium. The stress intensity factor at the top
(K/,top) calculated from the net pressure distribution should equal the fracture tough-
and
ness of the upper layer (K{C,2), and the stress intensity factor at the bottom (KI.boltom)
calculated from the same net pressure distribution should equal the corresponding
material property of the lower layer, (K{C.3)'
f b[Y, 4J, kI] = J (4J + ~ klY)
-Y
--dy.
l+y
(11.6)

For the case where the density is considered zero, Warpinski and Smith [5] give Selecting suitable integration constants, the integration results in
the following very elegant and concise system of equations:

V(l+Y
1-Y x [-ko (2ko+k1)y kly2]
..Jii(K{C.2 + K{C,3)
2v~
r;: = (
(T2 - (Tl
).
arCSIn -
(bc2)
ft[Y, ko, kI] = - kl + 2 +2

+ (2ko 2+ k1) arctan (Yv'(l + y)/(1 - y)


+ (0"3 - (Tl) arcsin (hp ~ b 2) - (T2 + 0"3 - 2p)i (11.1) l+y
, (11.7)

and
and
/b[y, ko, kI] = V{l=Y
1+Y x [ko - kl +
(24J - k1)y
2 +2
kly2]

-
(2ko - k1)
arctan
(Yv'(I- y)/(l + y) . (11.8)
There are only two unknown variables (c and b2) if the pressure, p, is given. Any 2 -1 + Y
root-finding numerical method can be used to obtain the solution.
However, here we are interested in the more realistic case where the density of The following limits are needed:
the fluid is not zero. In the following we derive the system of equations for the more Jr .
general case. fl[+l,ko,kI] = -(+2ko+kl),
4
The stress intensity factor for asymmetric loading is given by Eq. 2.55, which n
has to be applied in the vertical direction. Eq. 2.55 is rewritten in terms of the fl[ -1, 4J, kI] = -( -2ko - kJ),
. 4
dimensionless variable, Y, as n
10[+1, 4J, kI] = 4 (+24J - kt), and

.__ h_.p
__
Jr(Yu-Yd)
x e
L,
pn(y)J_1_+_Ydy
1-y
(11.3) (11.9)
274 Fracture height growth Equilibrium fracture height 275

From Figure 11.2 it is obvious that the integral in Eq. 11.3 consists of three parts Example 11.1 Equilibrium Height Neglecting the Density of the Fluid
corresponding to the intervals (1) from -1to Yd, (2) from Yd to Yu, and (3) from Yu
to 1. The slope, k1, of the net pressure is identical in these intervals and it is given Calculate the upper height growth and lower height growth for the selected data
in Table 11.1. The constant term, ko varies from interval to interval. For instance, in sets given in Table 11.3. (Note that in Warpinski and Smith [5J, the inverse problem
is considered to determine the pressure providing exactly 30 ft [9.16 m] of height
the interval from -1 to Yd, it is given by koo - 0"3, where koo is the pressure at the
migration.)
middl~ o~ the crack (note that it is not the pressure at the middle of the perforation).
Substituting the values of ko into Eq. 11.3 provides the expression for the stress
intensity factor at the top: Solution
The system of Eqs. 1L 12 and 11.13 are solved using the program of Valk6 and Econo-
X {I,[Yd, koo - 0"3. k1] - I I[ -1. koo - 0"3. k1] mides (12]. The results are presented in Table 11.4. Though not justified by the likely
it ( Yu - Yd )
accuracy of the input data, we use four significant digits to avoid further misunder-
+ I,[yu, koo - 0"1. kd - I,[Yd, koo - O"J, kl] + 11[+1, koo - 0"2, kl] standing. After the solution is obtained, we substitute it into Eqs. 11.1 and 11.2 to
check consistency.
(11.10)
The solution for the first case agrees with Warpinski and Smith [5] very well, with
where the variables koo and kl are given in Table 11.1 in terms of the two key an upper height growth of 30 ft calculated for the specified pressure. (In Warpinski
and Smith [5] the pressure was calculated from the height growth, but the calculations
variables, Yu and Yd. The function It is defined by Eq. 11.7, and the two limit
should result in the same consistent data set either way.) Moreover, Eqs. 11.1 and 11.2
values needed are given by Eq. 11.9. For the stress intensity factor at the bottom a
similar expression is obtained,
Table 11.3 Input data for Example 11.1 (After Simonson et al. [4])
hp No (Tl 0) K1C.2 K1C,) Ppc
K',bottom(Yu, }a) = ( ) X {!b[Yd, koo - 0"3, kl] - !b[-I, koo - 0"3.kl]
7r Yu - Yd
1 3500 psi 3500 psi 0 0 3350 psi
+ h[yu. koo - 0"1, k1] - !b[Yd. koo - 0"1>kl] + !b[+I, koo - 0"2, kl] 24.13 MPa 24.13 MPa 23.10 MPa
2 3500 psi 3500 psi 1000 psi inl/2 1000 psi- in 1/2 3360 psi
- !b[y", koo - 0"2. kl]}, (11.11) 24.13 MPa 24.13 MPa LOI MPa ml/2 1.01 MPa ml/2 23.17 MPa
3 3500 psi 4000 psi 1000 psi in1/2 1000 psi- inl/2 3360 psi
where all the functions and variables are defined as before.
24.13 MPa 27.58 MPa 1.01 MPa ml!2 1.01 MPa ml/2 23.17 MPa
The system of equations to solve is simply
4 3500 psi 4000 psi 4000 psi- inl/2 1000 psi- inl/2 3360 psi
(11.12) 24.13 MPa 27.58 MPa 4.04 MPa ml/2 1.01 MPa m 1/2 23.17 MPa

p=o 01 ::: 3000 psi (20.68 MPa) hp = 50 ft (15.24 rn)


,~

.~. and
(11.13)
Table 11.4 Solution of Eqs. 11.12 and 11.13 for the data sets of Table 11.2
Once the input data hp, 0"1, 0"2, 0"3, K1C.2, K1C,3, and Pcp are given, Eqs, 11.12 and No f)..h. f)..hd Eq. 11.1 Eq. 11.1 Eq. 11.2 Eq.IL2
11.13 contain only two unknowns, namely Yu and Yd. If another set of key variables left right left right
is preferred, Table 11.2 can be used to obtain a system containing only the pair Sh; (MPa) (MPa) (MPa) m (MPa) -m
and t:,.hd, or the pair b2 and c. Using a suitable root finding method, the solution can
30.07 ft 30.07 ft 0 0 0 0
be easily found.
9.164 m 9.164 m
The above relationships were obtained and used with the algebraic manipulation
2 26.08 ft 26.08 ft 0.4936 0.4936 0 0
system Mathematica (Wolfram [11]); a program was published in Valko and Econo-
7.950 m 7.950 rn
mides [12]. In the following we consider a series of examples extracted from pages
3 22.75 ft 4.241 ft 0.5686 0.5686 0 0
70 and 71 of Warpinski and Smith [5] and compare our results with the ones given
6.934 m 1.293 m
there. Once the solution corresponding to the simpler data set (with zero density) is
4 9.239 ft 4.051 ft 1.568 1.568 -2.921 -2.921
found satisfactorily, we can proceed to examine the more complex problem involving
2.816 m 1.235 m
hydrostatic pressure effects.
276 Fracture height growth Equilibrium fracture height 277

are satisfied by our solution perfectly as seen from Table 11.3; here we present both Table 11.6 Solution of Eqs, 11.12 and 11.13 for the data set
sides of Eqs. 11.1 and 11.2 as calculated at the upper and lower height growth shown of Table 11.4
in the same row of the table. First solution Second solution
The second case raises some doubts because while Warpinski and Smith [5] calcu- Pcp - 0"1
lates both the upper and lower height growth to be 30 ft, our value is only 26 ft.
However, because our solution satisfies exactly Eqs. 11.1 and 11.2, we conclude that llh. llhd llhu llhd
the discrepancy is not conceptual.
The third case shows a larger discrepancy between the numerical results of our 3 MPa 13.1 m 2.32 m
400 psi 29.9 ft 5.9 ft
calculations and those of Simonson et al. [4] where 30 it is calculated for the upper
5 MPa 208 m 34.8 m 197 m 238 m
height growth and 3 ft for the lower. The reason for this discrepancy seems to be that
700 psi 602 ft 84.3 ft 521 ft 889 ft
the fracture toughness was neglected by them. In this example, neglecting the fracture
toughness has a large impact on the calculated height.
The fourth case is not considered in Warpinski and Smith [5]. It is selected here to
illustrate non-identical critical stress intensity factors. 0 is provided in field units. Two possible height pair solutions are shown for 5 MPa to
illustrate multiplicity, and a similar case with 700 psi net pressure is provided in field
units. The results are given in Table 11.6.
11.1.4 The Effect of HydrostatiC Pressure The "height map" of the system, i.e. the location of the upper and lower tips vs.
the treating pressure, is shown in Figure 11.3. The location of the second solution is
From now on we do not neglect the hydrostatic pressure gradient in the fluid. Because indicated in the figure by dashed lines. At lower net pressures the second solution is very
of the hydrostatic term, the system of non-linear equations, i.e. Eqs. 11.12 and 11.13, far from the known solution and might explain why it has been neglected. However,
becomes more complicated and the problem of multiple solutions cannot be excluded.
The issue is illustrated by a numerical example which is essentially case No. 3
as the treating pressure increases, the two pairs of solutions converge. In this case,
there is no justification in neglecting the second solution-pair. There is only one pair
of Example 11.1, except that the density of the fracturing fluid is not taken to of solutions below the net pressure of 3660 psi (25.2 MPa). From 3670 to 3780 psi
be zero.

Example 11.2 Equilibrium Height Taking into Account the Density of Tip location Tip location
the Fluid [ft] 1m]
1000 300
Calculate the upper and lower height growth from the data given in Table 11.5. Investi- L_
800
gate the range of pressure which results in a unique solution-pair (i.e. upper and lower
V 200
height migration), two possible solution-pairs, and no solution. 600
400
v I
-:
........... 100
Solution
The system of Eqs. 11.12 and 11.13 is solved using the program of Valko and Econo-
mides [12]. In order to reveal possible multiplicity, different starting points are used
200
o
-200
- V
-r-....
\
o

-100
-400
in the solution algorithm. The solution of the system is shown for net pressure 3 MPa
to illustrate the case of a unique solution. A similar case with 400 psi net pressure .s00 I
/ -200
-800
./
/
Table 11.5 Input data for Example 11.2 -1000 -300

hp 50 ft 15.24 m -1200
0"1 3000 psi 20.68 MPa 3000 3100 3200 3300 3400 3500 3600 3700 3800 psi
O"z 3500 psi 24.13 MPa I
0"3 4000 psi 27.58 MPa 21
K1C2 1000 psi inl/2 1.01 MPa ml/2 Treating pressure
K1C3 1000 psi- inl/2 1.01 MPa mliZ
p 62.4 lbm/ft" 1000 kg/rrr' Figure 11.3 Height map for Example 11.2. The first solution pair of the height migrat!on
is shown with sclidIine. The dashed line corresponds to the second solution
278 Fracture height growth Three-dimensional models 279

(25.3 to 26 MPa), there exist two solution-pairs.Above 3780 psi (26 MPa), there is that it evolves through equilibrium states. Again, the reverse application of super-
no solution. The maximum equilibrium heigh! is approximately 1400 it (430 m). No position seems to be necessary to use net pressure calculated with varying (J"H.min,
equilibriumcan exist beyond this limit. 0 i.e. with varying boundary condition at infinity. To our knowledge, all the three-
dimensional models of hydraulic fracturing assume that the net pressure (calculated
In the previous example (and in every other three-layer equilibrium problem), the
as the difference between the pressure inside the fracture and the varying (J"H.min) can
second solution appears because the system is coupled. Downward height migration
be used to determine stresses and displacements as if this net pressure were acting
primarily decreases the stress intensity factor, because the net pressure in the lower
inside the fracture and the far-field stress state were identically zero.
layer is primarily negative and adds a negative term to the integrals in Eqs. 11.3
and 11.4. However, hydrostatic pressure has a secondary effect of increasing the net
pressure at the bottom. The larger the downward height migration, the larger is this 11.2.1 Surface Integral Method
secondary effect. Here a second solution may appear.
For the sake of simplicity, we consider only constant minimum stress inside a Clifton [13] refers to Bui [14] when stating that the change in normal stress on the
given layer, but the developed mathematical apparatus can be readily applied to stress crack plane is related to the crack opening, w(x, y), by an integral of the form

J1V'w
profiles varying linearly in a layer with different coefficients from layer to layer.
The second solution is not stable in the sense that the downward tip "runs away" ts pt; y) =: p(x, y) - (J"~(x, y, 0) = s, V'(l/R)dA', (11.14)
for any positive perturbation of the downward height migration. (In the case of
non-zero stress gradients in the formation, the non-stable tip might be the upper tip where A is the fracture surface, Ee is the effective elastic modulus
depending on the relative magnitude of the hydrostatic pressure gradient and stress
gradient.) G
Ee = , (11.15)
It is well known that there is an upper net pressure limit for which an equilibrium 4:rr(1 - v)
height exists. The important additional messages of Figure 11.3 are that (1) the
equilibrium height is not unique above a certain net pressure, which is considerably i.e. E. = E'/(8:rr) with the notation used in this book; V is the gradient operator
less than the "run-away pressure", and (2) there is a maximum equilibrium height
which can be created without a run-away. - a-
"1;;: -i+-j
0- (11.16)
In the petroleum engineering literature, it has been assumed that a gradually ox' oy"
increasing pressure establishes a (large enough) equilibrium height. Height growth, and the distance R between point (x', y') at which the integrand is evaluated and
in tum, limits the pressure rise. In other words, equilibrium height theory (with
point (x, y) at which the pressure is evaluated is
incomplete numerical formulas) has suggested that the system is self-regulating. The
fact that there is a limit to the height which can be achieved in a stable manner R = [(x - x? + (x - x?]1/2. (11.17)
creates a conceptual problem; i.e. the self-regulating mechanism is mostly an artifact
of neglecting gravity. In the original paper by Bui [14], the problem of a plane crack under pressure
For problems with more than three layers, several solution-pairs may be possible acting on its faces but without any far field stress is considered. Introduction of
for a given treating pressure and more than one solution pair may be stable. a~(x, y, 0) (the value of the normal compressive stress on the crack surface at time
The equilibrium fracture height concept can be used to determine a realistic height =
t 0) into the definition of net pressure by clifton et al. [15] is an attempt to avoid
for a constant height model. The general procedure is to assume a constant height, the use of the concept of varying far field stresses.
run a simulation, determine a representative treating pressure, obtain the equilibrium Both the crack opening and the pressure are represented as linear combinations
height corresponding to the representative treating pressure and repeat the constant of suitable selected trial functions [13]. To construct the trial functions, the crack
height simulation with the improved estimate of fracture height. Such a procedure is surface is covered by a quadrilateral mesh as shown in Figure 11.4. The i-th trial
suggested in Rahim and Holditch [1]. function has the value of unity at the i-th node, varies linearly over adjacent triangles,
If the height is to be calculated simultaneously with the other two characteristic and vanishes along the opposite side of these triangles. Essentially the same trial
dimensions (length and width), then three-dimensional models are needed. functions are used for the crack opening and the pressure, except for the special
elements introduced in the near-crack-tip zone for the crack opening. (These special
trial functions vary as the square root of distance from the tip.) Substituting the
11.2 Three-dimensional Models representation of the pressure and crack opening results in a large (and dense) set of
The equilibrium height theory in Section 11.1 was presented for a vertical line crack. linear equations:
A natural generalization is to consider a fully three-dimensional fracture and assume Kw = T.6.p, (lLl8)
._- ---.-- .-_. --_ ... _----------------------

280 Fracture height growth Three-dimensional models 281


y
and
Wellbore element
ap + I (!1!)n'-l qy = F (11.21)
aY 'I w
2
w
3 P y,

lip element where the product pF y is the body force per unit volume caused by the weight of
the fluid. The "viscosity parameter", 'I', is given by

ri' :::: in'+l)K' (2 + :,) n' , (11.22)


-- x
-:.">-:".:
_- with K' and n' being the "usual power law coefficients". At sufficiently low values
of shearing rates (jqjlwZ), the power law fluid is replaced by a Newtonian fluid to
avoid singular coefficients in Eqs. 11.21 and 11.22_
Equations 11.20 to 11.22 are not used directly but are reformulated using a vari-
ational approach. The boundary condition at the moving boundary is derived from
the requirement that the flow into the near-crack-tip region is balanced by the sum
of three phenomena: leakoff rate, spurt loss and local fracture volume expansion.
Advance of the crack occurs in such a way that the stress-intensity factor, K/, is
Figure 11.4 Surface boundary elements used in one of the three-dimensional fracture propagation kept nearly equal to the critical stress-intensity factor, K{c, during crack extension
models, after [13] at each node. In principle, the stress intensity factor can be calculated once the
nodal crack openings are known for the given time step. For technical reasons, the
where w is the vector of nodal displacements, K is the stiffness matrix, T is the models may avoid calculating the stress intensity factor explicitly but make use of
matrix of element areas and !!..p is the vector of nodal net pressures. The derivation an equivalent equation to fulfill the equilibrium requirement.
of the mat~x coefficients in K and T is presented by Clifton and Abou-Sayed [15] The three-dimensional model described above is based on the displacement discon-
where t~e interested reader may find further details. The main advantage of the tinuity method, which is only one of the many available numerical methods in linear
elasticity _Any of the methods belonging to thelarge arsenal of numerical mechanics
surface mtegral method (or in other words, the boundary integral method) is that
can be selected as a starting point for a three-dimensional hydraulic fracturing model.
the stiffness matrix has less rows (and columns) than the similar matrix of the finite
Finite elements, discrete elementsand their numerous variations are potential candi-
element method. On the other hand, it is dense while in the finite element methods
it is sparse, dates [17-19]_ The fluid flow description submodel might also utilize one of these
methods, In general, the overall behavior of the model should not depend signifi-
Another large set of equations involving the same variables is obtained from the
cantly on the selection of the numerical method. The crack advance criterion and its
two-dimensional continuity equation:
practical implementation. should be more influential.

2CLO(p - Plr) aw
-r===;:::::::::=';:-'- - - +ai (11.19)
./t - rex, y) at ' 11.2.2 The Stress Intensity Factor Paradox

where qx is the volume flow rate in the x direction per unit length in the y direction Clifton [13] presents a series of case studies. In his example Case A (Newtonian fluid,
and_qy is the volume flow rate in the y direction per unit length in the x direction; no leakoff, high stress barrier) the fracture propagates with essentially constant height
q{ IS the volume injection rate per unit fracture area (non-zero only in the near- after the first few seconds, The important result is that the treating pressure starts
wellbore elements); the product CL,O x (p - Plr) is the Carter leakoff coefficient to increase after a short time and "the three-dimensional predictions of width and
(now depending linearly on the difference of the pressure inside the fracture and pressure agree quite well with the predictions of the two-dimensional PKN model",
the fluid pore pressure in the reservoir, pfr); and rex, y) is the opening time of the (It is interesting to note that Barree [16] arrived at a similar conclusion, finding that
a three-dimensional model behaves similarly to a KGD model at early times and to
surface element located at the position (x, y). Two pressure gradient equations are
included: a PKN model at late times.)
Simple reasoning suggests that a three-dimensional model based on the frac-
ap + ri'
ax
(lil)n'-l qx
w2 w3
= 0, (ll.20) ture toughness criterion would give decreasing treating pressures with time, so the
increasing treating pressure might initially seem a bit surprising, Let us assume
282 Fracture height growth
Pseudo-three-dimensional models 283

temporarily that the stress intensity factor at the lateral tip can be estimated as ihe
product of the square root of half-length and some kind of average net pressure. 1
Since for a well contained fracture (e.g. Case A of Clifton [13]) the length and the ~
pressure are increasing simultaneously, it is difficult to see how the stress intensity
factor remains "nearly constant". The apparent contradiction between the behavior of
-
<,

N
.......
....
0.8

the three-dimensional model and the intuitive comprehension of the stress intensity ~
'-'
0.6
factor can be resolved in two ways. .3.
.......
One possibility is that the net pressure is decreasing with time in average while the ~
0 0.4
value at the wellbore is still increasing. According to our understanding in that case
~
the pressure distribution in the fracture would become steeper and steeper near the a:
wellbore with increasing time. (However, it is difficult to see a mechanism providing 0.2
larger pressure drops near the wellbore with increasing time, knowing that also the
width is increasing and the flow rate is nearly constant in this region.) 0
The other possibility is that the stress intensity factor at the lateral tip is very a 2 4 6 8 10
much influenced by the fact that the aspect ratio, (2xf / hf) is increasing with time.
To understand the phenomenon, we substitute the fracture with an idealized three- Aspect ratio, 2x,/ h,
dimensional elliptical crack with constant height, constant inner pressure and variable Figure 11.6 Normalized stress intensity factor at the tip of the pressurized plane crack of ellipt~cal
half-length. Such a crack is shown in Figure 11.5. The stress intensity factor at the shape (curve a). For comparison the normalized stress intensity factor at the vertical
elliptical boundary is given analytically by Kassir and Sih [20] (see page 82 of their tip is shown as curve b. Curve c is the function (2xtI h f )-1/2
book). In our notation,
The stress intensity factor at the lateral tip, normalized by pVhfl2 is shown
(11.23) as curve a of Figure 11.6. As seen from the figure, when the aspect ratio (~f / ~f)
reaches the value five, the normalized stress intensity factor becomes almost Identical
to the function (2x tIhf) -1/2, also shown in the figur~ as c~rve c. Thus, t?e ch~ge
where the k parameter depends only on the aspect ratio,
in the aspect ratio Significantly decreases the stress intensity factor and. mcreasl~g
treatment pressure is needed to compensate this effect. From the asymptotic behavior
(11.24) of the stress intensity factor, it follows that in a three-dimensional model of a well-
contained hydraulic fracture the pressure increases with the square root of length
cp is the angle with the x axis (it is zero for the lateral tip stress intensity factor at late times. (Eq. 9.11 shows a somewhat different result obtained from the PKN
shown in Figure 11.6 and ](/2 for the stress intensity factor at the vertical tip) and width equation where the pressure increases with the fourth root of l~ngth.) .
E(k) is the complete elliptic integral of the second kind. While the analytical solution for the stress intensity factor at the tip of an elhp-
tical crack seems to explain the "stress intensity factor paradox", it is not quite
clear whether other idealized geometries, e.g. a rectangular fracture with circular
propagating edges, would also result in a decreasing stress intensity factor at the
lateral tip. . .
For comparison, the normalized stress intensity factor at the vertical ttp of ~he
elliptical crack is also shown in Figure 11.6 (as curve b). At lar~er aspect ratl~s
it converges to unity, i.e. to the value anticipated from the vert.ICal plane st~am
condition. (This illustrates that for the equilibrium height calculations, the vertical
line crack approximation is reasonable.)
1----- x, --- ~I
.....
11.3 Pseudo-three-dimensional Models
Figure 11.5 Pressurized plane crack of elliptical shape. The inner pressure p and the height is The pseudo-three-dimensional (P-3D) concept of mode!ing ~y~raulic fract.uring was
constant, the length is varying
introduced by Settari and Cleary [2]. A P-3D model IS a limited extension of the
284 Fracture height growth References 285

differential two-dimensional model (in most cases that of Nordgren [21]), allowing 4. Simonson, E.R., Abou-Sayed, AS. and Clifton, RJ.: Containment of Massive Hydraulic
for variable height. The height as a function of the lateral position is calculated in Fractures, SPEJ, (Feb.), 27-32, 1978.
every time step. Height growth is controlled by the pressure which is often consid- 5. Warpinski, N.R. and Smith, M.B.: Rock Mechanics and Fracture Geometry, in Recent
ered to be a function of the lateral position only. For the determination of height, Advances in Hydraulic Fracturing, J. Gidley et al. (eds.): Monograph Series, SPE,
every vertical cross section is considered separately; the resultant height-growth rates Richardson, TX (1989), SPE, Richardson, TX, 1989.
6. Ahmed, U.: Fracture-Height Predictions and Post-Treatment measurements, in Reservoir
(Settari [221) or the resultant height-growth values (Palmer and Caroll [23]) are then
Stimulation (2nd ed.), Econornides, M.J. and Nolte, K.G. (ed.) Prentice Hall, Englewood
used in the continuity equation.
Cliffs, N.J., 1989.
The concept was used by Nolte [24] and Palmer and Luiskutty (25] and subse- 7. Newberry, B.M. Nelson, R.F. and Ahmed, U.: Prediction of Verital Hydraulic Fracture
quently by many other authors. In principle, the methods of Section 11.1 can be Migration Using Compressional and Shear Wave Slowness, Paper SPE 13895, 1986.
applied at any lateral cross section to obtain the equilibrium height from the actual 8. van Ekelen, H.AM.; Hydraulic Fracture Geometry: Fracture Containment in Layered
pressure. The equilibrium height (if it exists) is considered as an upper bound Formations, SPEJ, (June), 341-349, 1982.
and some models attempt to limit the height growth by other factors, e.g. fluid 9. Fung, R.L., Vijayakumar, S. and Cormack, D.E.: Calculation of Vertical Fracture
flow [26-28]. If the rate of height growth is limited by the energy dissipation Containment in layered Formations, SPE FormationEvaluation, (Dec.), 518-522, 1987.
due to creating a damaged zone, a continuum damage mechanics approach might 10. Economides, MJ.: A Practical Companion to Reservoir Stimulation, Elsevier,
be useful [12]. A detailed discussion of the P-3D or 3D computer models actually Amsterdam, 1992.
used in the petroleum industry is beyond the scope of this book as these computer 11. Wolfram, S.: Mathematica:A System for Doing Mathematics by Computer, 2nd ed.
programs themselves are sophisticated systems with numerous additional algorithmic Addison-Wesley, Reading, MA 1991.
12. Valko, P. and Economides, M.J.: Fracture Height Containment with Continuum Damage
rules. These rules reflect the aggregate experience of modelers involved in everyday
Mechanics, Paper SPE 26598 presented at the Annual Technical Meeting and Exhibition
field jobs for many years and are extremely useful in bridging the gap between theory
Houston, 1993.
and practice.
13. Clifton, R.J.: Three-dimensional Fracture propagation Models, in Recent Advances in
In the past, a certain hierarchy of models was accepted. Two-dimensional Hydraulic Fracturing, (J. Gidley et al. (eds.j), Monograph Series, SPE, Richardson,
models have been considered as the routine vehicle for everyday design and TX,1989.
analysis of fracturing jobs. Pseudo-three-dimensional models have been used as 14. Bui, H.D.: An Integral Equations Method for Solving the Problem of Plane Crack of
improved engineering tools needing more accurate input data and somewhat increased Arbitrary Shape, J. Mech. Phys. Solids, 25, 29-39, 1977.
computational time. Fully three-dimensional models have been thought of mainly as 15. Clifton R.J. and Abou-Sayed AS.: On the Computation of the Three-dimensional Geom-
computationally demanding research tools which need even more input information etry of Hydraulic Fractures, Paper SPE 7943, Symp. on Low-Permeability Reservoirs,
but provide a deeper understanding of fracture propagation. With the development Denver, 1979.
of computer hardware and software, this strict hierarchy of models is becoming 16. Barree, R.D.: A Practical Numerical Simulator for Three-dimensional Fracture Propaga-
obsolete and the emphasis is shifted toward understanding the coupling of various tion in Heterogeneous Media, Paper SPE 12273 presented at the Reservoir Simulation
Symposium, San Francisco, Nov. 15-18, 1983.
subprocesses. The inclusion of previously overlooked (but lately proven effects) may
17. Advani, S. H.: Finite Element Model Simulations Associated With Hydraulic Fracturing,
improve the prediction of fracture geometry and propagation even more significantly
SPEJ, (April), 209-218, 1982.
than the attempt to incorporate the latest numerical methods found promising in other 18. Thiercelin, MJ, Ben-Naceur, K and Lernanczyk, Z.R.: Simulation of Three-
fields of mechanics. dimensional Propagation of a Vertical Hydraulic Fracture, Paper SPE 13861 presented
at the Low-Permeability Gas Reservoir Symposium, Denver, May 19-22, 1985.
19. Morita, N., Whitfill, D.L. and Wahl, H.A: Stress Intensity Factor and Cross Sectional
References Shape predictions from 3D Model for Hydraulically induced Fractures, IPT, (Oct.),
1329-1342, 1987.
1. Rahim, Z. and Holditch, S.A: Using a Three-dimensional Concept in a Two-dimensional 20. Kassir, M.K. and Sih, G.C.: Three-dimensional crack problems, in Mechanics of Frac-
Model to Predict Accurate Hydraulic Fracture Dimensions, Paper SPE 26926 presented ture, Vol. 2, Sih, G.C.(ed.), Noordhoff, Leyden, 1975.
at the Eastern Regional Conference and Exhibition, Pittsburgh, Nov. 2-4, 1993. 21. Nordgren, R. P.: Propagation of a Vertical Hydraulic Fracture, SPEJ, (Aug.), 306-314,
2. Settari, A. and Cleary, M. P.: Development and Testing of a Pseudo-Three-dimensional 1972; Trans.AlME, 253.
Model of Hydraulic Fracture Geometry, SPEPE, (Nov.), 449-466, 1986; Trans.AlME, 22. Setrari, A: Quantitative Analysis. of Factors Influencing Vertical and Lateral Fracture
283,1986. Growth, SPE ProductionEngineering, (Aug.), 310-322, 1988.
3. Bouteca, M.J.: Hydraulic Fracturing Model Based on a Three-dimensional Closed Form: 23. Palmer, J.D. and Caroll, H.B. Jr.: Numerical Solution for Height of Elongated Hydraulic
Tests and Analysis of Fracture Geometry and Containment, SPE ProductionEngineering, Fractures, Paper SPE 11627 presented at the Low Permeability Gas Reservoirs Sympo-
(Nov.), 445-454, Trans.AlME, 285, 1988. sium, Denver, March 13-16, 1983.
_._--_._----_._._._----------------

286 Fracture height growth

24. Nolte, K.G.: Principlesof Fracture Design Based on Pressure Analysis SPEPE (F b)
22-30, 1988. ' ,e .,
25. Palmer, LD. a~d Luiskutty,C.T.: A Model of the Hydraulic Fracturing Process for
Elongated Vertical Fractures and Comparisons of Results with Other Models, Paper
SPE 13864,presentedat the Low Permeability Gas Reservoirs Symposium, Denver
May, 1985. ,
26. Moral~s,R'!1' and Abu-Sayed,A.S.: Microcomputer Analysis of Hydraulic Fracture
Behavior with a Pseudo-Three-dimensionalModel, SPEPE, (Feb.), 198-205, 1989.
27. ~eyer, ~.R, Cooper,G.D. and Nelson, S.G.: Real-Time 3-D Hydraulic Fracturing
Slmul~llon;Theory and Field Case Studies, Paper SPE 20704 presented at the Annual
Technical Conf. and Exhibition,New Orleans, 1990.
28.. Weng,X.: Incorporationof 2D FluidFlow into a Pseudo-3DHydraulicFracturingSimu- APPENDIX: COMPARISON
lator, Pap~r SPE 21849 presented at the Rocky Mountain Regional meeting and Low
PermeabilityReservoir Symposium,Denver, Co, April 15-17, 1991. STUDY OF HYDRAULIC
FRACTURING MODELS:
INPUT DATA AND RESULTS

Warpinski et al. (1] is a summary of a comparative study of hydraulic fracturing


models using test data from the GRI Staged Field Experiment No.3. More details
of the SFE3 experiments are given in Refs. 2-5.
Models compared in the study include 2D, pseudo-3D, and 3D codes, run on up to
eight different cases. The purpose of the study was "to provide production engineers
with a practical comparison of the available models so that rational decisions can be
made as to which model is optimal for a given application".
Table Al shows the relevant rock and reservoir information for the comparative
study. Three different physical configurations were considered: single-layer (2D)
case, three-layer (3D) case and five-layer (3D) case. Table A2 gives the character-
istics of a treatment. The participants were asked to model the treatment. The tables
are reproduced from Warpinski et al [1]. We added the symbols as used in this book
whenever they were available.
Each participant could model a total of eight cases. These were GDK (KGD),
PKN, three-layer and five-layer cases, with separate runs for a constant Newtonian
viscosity and a constant n' and K' power-law fluid. The PKN and GDK cases were
run with a constant height (2D) set at 51.8 m (170 ft). The three- and five-layer cases
were run with a 3D or pseudo-3D model, allowing fracture height be determined by
the model.
The results provided by the modelers are summarized in Tables A3-AS
Tables A6-A8 show those names which were used to identify the computer codes
(model runs) by Warpinski et al. [1]. An interesting discussion of the study was
published in the same issue of the journal (Cleary [6]).
---~-------- ------------~---------------------- -----

288 Appendix Appendix 289

Table Al Rock and reservoir data Table A.3 2D-Results at end of pump
Inter- Depth zone III situ poisson's young's Fracture Length Height Pressure Maximum width Average Overall Efficiency
val (ft) thickness stress ratio modulus toughness (ft) (ft) (psi) (in) width at average (%)
(ft) (psi) (lQ6 psi) (psi/in 1/2) weUbore fracture
min max (in) width
v E
(in)
Single layer (2D) case
pn,w (*)
1 9170 9340 170 5700 0.21 8.5 2000
200 cp
Three-layer (3D) case
2542 170 62 0.848 0.849 0.605 85.5
1 8990 9170 180 7150 0.30 6.5 2000 2 4855 170 1094 0.502 0.397 0.289 72.3
2 9170 9340 170 5700 0.21 8.5 2000 3 2584 204 1685 0.91 0.76 0.73 93
3 9340 9650 310 7350 0.29' 5.5 2000 4 2659 170 70 0.79 0.79 0.62 83.1
Five-layer (3D) case 5 4507 170 1188 0.55 0.43 0.32 72.2
6 2288 170 97 0.94 0.94 0.74 85.4
1 8990 9170 180 7150 0.4 76.6
0.30 6.5 2000 7 3803 170 1474 0.68 0.53
2 9170 9340 170 5700 0.21 8 2724 170 53 0.78 0.78 0.61 84
8.5 ~OOO
3 9340 9380 40 7350 0.26 5.4 2000 9 4039 170 1377 0.59 0.46 0.37 75
4 9380 9455 75 5800 0.20 7.9 2000 10 2480 200 71 0.74 86
5 9455 9650 195 8200 0.30 4.0 2000 11 4157 200 925 0.50 77
12 1347 170 81.9 0.77 0.77 0.6 81.9
13 2029 170 1380 0.63 0.36 73
Table A2 Treatment data 14 4595 170 1182 0.54 0.43 0.32 73.8
15 2212 170 82 0.98 0.98 0.77 '85.9
Bottomhole temperature, 'F 246 16 2716 170 0.767 0.6 82.5
Reservoir pressure, psi p, 3600 17 3986 170 0.554 0.37 74.4
Spurt loss s, 0.0 18 3866 170 1595 0.627 0.492 0.387 75
Fluid leakoff height Entire fracture height 19 3556 170 1684 0.704 0.553 0.434 75
Fluid leakoff coefficient, ftlmin 1/2 CL 0.00025
Viscosiry - case A, cp f.1. 200
Rheology - case B
n' 20 2542 170 6l.8 0.85 0.85 0.6 61.8
0.5 21 4629 170 .1067.5 0.54 0.42 0.28 73.6
K', Ibfso.5IW (*) 0.06 22 2516 204 1624 0.98 0.82 0.75 93
Injection rate, bpm 50 23 2098 170 117 1.04 1.04 0.82 86.4
Fluid Volume, bbi 2Vi 10000 24 4118 170 1397 0.64 0.5 0.36 74.3
Proppant None 25 1808 170 161 1.24 1.24 0.97 88.3
Assumed unit, no! given in Warpillski et al. [1] 26 3395 170 1774 0.81 0.64 0.46 79
27 2142 170 89 1.03 1.03 0.81 89
28 3347 170 1754 0.75 0.59 0.47 79
29 4046 170 1474 0.68 0.53 0.38 76.9
30 2031 170 97 1.07 1.07 0.84 86
31 2304 170 0.933 0.933 0.733 85.2
32 3656 170 0.622 0.415 76.5
33 3396 170 1880 0.738 0.58 0.456 78
34 3155 170 1986 0.817 0.641 0.504 81.7

No symbol is used for this variable ill this book


--'_"-"'-- ,-- ---_.

290 Appendix
Appendix 291
TableA.4 Three-Layer results at end of pump
TabJeA,S Five-layer results at end of pump, Symbols given are notations used in this book
Length Height Pressure Maximum width Average Overall Efficiency Length Height Pressure Maximum width Average Overall Efficiency
(ft) (ft) (psi) (in) width at average (%) average
(ft) (ft) (psi) (in) width at (%)
wellbore fracture wellbore fracture
(in) width (in) width
(in) (in)
xf hj P".l\ WI1'.O (*) iii
xf hf P.,k' W"".O (0) w
200 cp
200 cp
3408 318 1009 0,65 0.35 0.3 77 1 2905 394 960 0.72 0.42 0.31 80.1
2 3750 903 283 0.56 0.32 0.25 66 2 3709 361 852 0.63 0.38 0.25 66
3 1744 544 1227 0.9 0.54 0.36 80 3 1754 501 1119 0.83 0.6 OA 82
4 l360 442 1387 L04 0.68 0.64 96 4 1224 476 1250 1.03 0.7 0.65 97
5 3549 291 987 0.58 0.35 0.29 70.3 5 2962 328 669 0.5 0.36 0.28 70.5
6 2697 360 1109 0.72 0.41 0.34 74.3 6 2407 327 768 0.6 0.46 0.35 74.8
7 3598 306 992 0.57 0.31 0.25 67 7 3399 394 944 0.64 0.36 0.24 68
8 1938 435 1132 0.72 68 8 2011 428 1008 0.68 69
9 2089 357 1113 0.66 0.33 0.25 43 9 1594 438 1129 0.81 0.45 0.36 58.1
.s: .s:
10 3259 371 1093 0.75 0.38 0.31 77.6 10 2647 430 1035.5 0.82 0.46 0.31 81.8
11 3289 329 1005 0.67 0,35 0.26 68 11 2765 388 935 0.71 0.42 0.25 0
12 902 596 1428 1.1 0.74 0.49 62 12 1042 600 1358 1.18 0.9 0.6 87
13 1326 442 1433 L08 0.71 0.66 96 13 1156 476 1262 1.04 0.71 0.66 93
14 2915 337 1094 0.69 0.4 0.32 72.7 14 2535 330 766 0.6 0.46 0.37 73.7
15 2120 413 1212 0.86 0.48 0.4 76.9 15 1980 349 891 0.75 0.57 0.42 77.8
16 3235 353 1083 0.65 0.33 0.26 69 16 2926 405 968 0.7 70
17 2424 435 1171 0.74 0.34 0.21 47 17 3124 449 1160 0.74 62
"No symbol is used for this variable in this book 18 1125 602 1270 1.11 76
19 2636 391 934 0.49 62
20 1870 458 1151 0.85 0.47 0.34 64
"No symbol for this variable is used in this book
Appendix 293
292 Appendix

Table A6 Model names corre- Table A.7 Model names corre-


sponding to entries in sponding to entries in
Table A.3 Table A.4
Model name as
Model name as
given in Warpinski et al. [1]
given in Warpinski et a/. [I]
1 SAH
1 SAH (GDK)
2 NSI
2 SAH (PKN) 3 RES
3 Marathon 4 Marathon
4 Meyer-l (GDK) 5 Meyer-I
5 Meyer-l (PKN) 6 Meyer-2
6 Meyer-2 (GDK) 7 Areo-Stimplan
7 Meyer-2 (PKN) 8 Texaco-FP
8 Shell (GDK) 9 Advani
9 Shell (PKN) 10 SAH
10 Texaco-FP (GDK) 11 NSI
11 Texaeo-FP (PKN) 12 RES
13 Marathon
12 Chevron (GDK)
14 Meyer-l
13 Chevron (PKN) 15 Meyer-2
14 Advani 16 Arco-Stimplan
15 Halliburton 17 Advani
16 Conoco (GDK)
17 Conoeo (PKN)
18 ENERFRAC-l Table A.S Model names corre-
sponding to entries in
19 ENERFRAC-2 Table A.5
20 SAH (GDK)
21 SAH (PKN) Model name as
22 Marathon given in Warpinski et al. [I]
23 Meyer-l (GDK) 1 SAH
24 Meyer-l (PKN) 2 NSI
25 Meyer-2 (GDK) 3 RES
26 Meyer-Z (PKN) 4 Marathon
27 Shell (GDK) 5 Meyer-l
28 Shell (PKN) 6 Meyer-2
7 Arco-Stimplan
29 Advani
8 Texaeo-FP
30 Halliburton
9 Advani
31 Conoco (GDK)
10 SAH
32 Conoeo (PKN) 11 NSI
33 ENERFRAC-l 12 RES
34 ENERFRAC-2 13 Marathon
14 Meyer-l
S. A. Holditch & Assocs.Inc.
15 Meyer-2
16 Arco-Stimplan
17 Arco- Terrafrac
18 Texaco-FP
19 Texaco-FPNOTIP
20 Advani
-
294 Appendix

References

1. Warpinski N.R., Moschovidis, Z.A., Parker C.D. and Abou-Sayed, 1.S.: Comparison
Study of Hydraulic Fracturing Models: Test Case - GRI-Staged Field Experiment No.
3, SPE Production & Facilities, 9 (1), 7-16, 1994.
2. Holditch, SA. et al.: The GRI Staged Field Experiment, SPE Production Engineering,
INDEX
(Sept.), 519, 1988.
3. Robinson, B.M. Holditch, S.A. and Peterson, R.E.: The GRI's Second Staged Field
Experiment: A Study of Hydraulic Fracturing, Paper SPE 21495 presented at the Gas
Technology Symposium, Houston, Jan. 22-24, 1991.
4. Robinson, B.M. et al.: Hydraulic Fracturing Reserarch in East Texas: The GRI Staged
after-growth 228, 230, 258 constant height model 267
Field Experiment, Journal of Petroleum Technology, (Jan.), 78, 1992. contained hydraulic fracture 267
5. Saunders, B.F. et al.: Hydraulic Fracturing Reserarch in the Frontier Formation through time of 260
anelastic strain recovery (ASR) 15,77 continuum damage mechanics (CDM) 240,
the GRf's Fourth Staged Field Experiment, Paper SPE 24845 presented at the Annual 247, 258
Technology Conference and Exhibition, Washington, Oct. 4-7, 1992. average width 166
convection 145
6. Cleary, M.P. Discussion of Comparison Study of Hydraulic Fracturing Models: Test bilinear flow 5 Couette flow 97
Case-GRF-Staged Field Experiment No.3, SPE Production & Facilities, 9 (1), 17-18, Bingham plastic model 99 crack layer theory 240
1994. BNS equation 133 crack opening 279
boundary element method 49 critical depth 58
boundary integral method 280 crosslinker 13
breakdown pressure 65, 68
damage
brittle material 237
accumulation 248
brittle solid 242
residual 11
Carter zone 247
equation I 169, 180 Darcy slaw 2
fluid loss model 183 Deborah number 136
leakoff 210, 212, 263 deviated well 90
equation II 172 differential strain curve analysis (DSCA) 77
II material balance 199 dilatancy 245
IIsolution 197 dilatant fluid 99
CDM version displacement 22
of the Nordgren-Kemp model (CDM-NK) dissipation 115
250 drag
of the Perkins-Kern-Nordgren model coefficient 139
(CDM-PKN) 252,260 reduction 134
characteristic dimension 216 drilling direction 71
characteristic length 105
circular crack 45, 47, 86 effective stress 55, 66, 71
energy of a 48 horizontal 73
closure pressure 13, 71, 73, 76, 79, 216, 263 elastic energy 47
closure quality 71 elastic region 26
Colebrook -White equation 131 equilibrium height 275
combined parameter 249, 252 equivalent Newtonian viscosity 110, 119,224
dimensionless 258
comparative study of hydraulic fracturing failure 237
models 287 far-field stress 55, 279
compressive strength 237 filter-cake 169
conductivity/porosity factor 71 resistance 184, 21i, 231
finite conductivity fracture 5, 232
confining stress 246
finite elements method 49, 242
consistency index 99
Index 297
296 Index
Rabinowitsch-Mooney equation 117
perforated interval 267 radial width equation 196
fissure total compressibility 186 injected volume 166
perforation phasing 61 radially propagating fracture 267
flow injection test 79
behavior index 99 ~nstantaneous shut-in pressure 74, 219 Perkins-Kern real-gas pseudopressure 2
curve 103, 109, 110, 115, 116, 118 mtaleyer stress contrast 79 model 83 relative roughness 131
no leakoff 191 reservoir pressure 55
laminar 97
Kachanov law 247 width equation 190, 192
turbulent 131 retardation 237
in annulus 122 Kachanov parameter 248 permeability 71 retarded tip propagation 226, 254
in elliptic cross section 123 KGD proppant-pack 11 Reynolds number 104, 108
in fracture 127 geometry 31, 85, 88, 214 fracture 8 generalized 133
in limiting elliptic cross section 124 model 203, 267 reservoir 15, 231 particle 139
flow-back test 76 Khristianovich and Zheltov model 83 PKN wall 135, 152
kinetic energy correction factor 104, 137, 153 geometry 31, 83, 88, 214
fluid rheological
efficiency 166 model 202, 267 constitutive equation 98
lag 245 leakoff 11,76 width equation 204 curve 98, 103, 110, 116, 118
leakoff 169 volume 165 plane strain 30 rotating cup viscometer 161
loss parameters 215 coefficient 258 horizontal 83, 85, 193 run-away pressure 278
foam 13 history of 187 modulus 263
quality 13, 147 LEFM 243 vertical 83,193 screenout 93
formal material balance 178 line crack 32, 238 plane stress 27 shape factor 85, 190, 192
frack & pack 8, 184 energy of a 48 plastic behavior 99 shear
fracture two-wing 47 plastic region 26 rate 97
coalescence 67 vertical 271 plug with uniform velocity 106, 113 nominal Newtonian 109
compliance 214 linear elastic fracture mechanics Poiseuille flow 97 stress 20, 54, 97
conductivity 15, 185 (LEFM) 240 poisson ratio 24, 44, 55, 75 skin effect 4, 15
dimensionless 5, 186 load-carrying capability 240 pore volume 2 slip
half length 167 longitudinal fracture 90 poroelastic constant 55, 75 apparent 158
height migration 92 porosity 2 coefficient 157
network area, dimensionless 186 matrix hydraulic diffusivity, dimensionless 186 power law correction 157
surface 166 matrix stimulation 4 assumption 173, 178 velocity 157
toughness 239, 249, 269, i81 maximum drag reduction asymptote (MDRA) generalized 133 smooth closing 38
apparent 246 134, 136, 152 length growth 201 Sneddon crack 87
fracturing fluid 11 maximum velocity 104, 107 model 99 specific
friction factor 134, 153 maximum width 34 width growth 180 surface energy 239
Fanning 104, 133 at the wellbore 85 pressure decline analysis 212, 215, 260 volume expansion ratio 148
Weisbach 104 minifrac test 217 Nolte's 212 spurt loss 169, 171, 256
minimum horizontal stress 58, 89, 92, 269 pressure transient testing 4 coefficient 217
g-plot 230, 257 Mooney plot 160 principal stress 21, 54 negative 221
Geertsma-de KIerk (GDK) moving boundary 245 superposition 44 steady-state 3
. width equation 193 process zone 246, 247 Stefan'S boundary condition 182
model, no leakoff 194 net pressure 48 propagation step-rate test 76
general fluid 117, 132 fracturing 76 criterion 242 Stokes' law 139
geometry factor 271 net present value (NPV) 7, 14 pressure 218, 220, 227
Griffith crack 32 strain 21
Nolte bounds 200 rate 208, 242 energy 47, 238
Griffith stability criterion 239 non-Newtonian behavior 202 velocity 235 relaxation 78
non-wetted zone 243 of longitudinal wave 25
Hagen-Poiseuille law 103 115 Nordgren equation 206 stress
proppant 13 absolute 56
height map 277 ' Nordgren-Kemp model 183, 209 carrying capacity 138
high net pressure 253 absolute vertical 56
numerical material balance 179 materials 8
Hook's law 214 anisotropy 72
settling 138 compressive 20
horizontal well 90 Oldroyd-Jastrzebski plot 160
transport 12 stress (continuetf)
breakdown pressure 69 opening time 181 pseudo-three-dimensional model 268
hydrostatic pressure gradient 276 distribution factor 171 contrast 14
pseudoplastic behavior 99 distribution 35, 42
overburden
pseudoradial flow 6 intensity factor 42, 43, 239, 240, 269, 281
induced stress 59 pressure 56
pseudosteady-state 3
infinite conductivity fracture 5, 72 stress, absolute 55
-
298 Index

nominal 248 turbulent flow


normalized 283
in ellipsoid cross section 137
in-situ 63
maximum principal 43 uniaxial test 25
minimum principal 43 unwetted zone at the tip 165
net section 248
nominal 248 velocity profile 132
normal 20 virial equation of state 153
singularity 35 viscosity 97
tectonic 56 apparent 99
tensile 20 apparent wall 110
wall 117 wall 119
stress-ahead function 248 volume equalized
constitutive equation 149
Tvshaped fracture 94 power law 151, 155
tensile failure 61 quantity 148
tensile strength 65, 237 shear rate 149
terminal settling velocity 140 yield-power Jaw 149
three-dimensional model 267 Bingham plastic 151
tip propagation velocity 249
tip retardation 7, 247 yield stress 99
tip screen-out 8 Young's modulus 23, 44, 55, 75
tortuousity 89, 227
transient leak-off 185 zero absolute pressure at the tip 195
transverse fracture 90 zero net pressure at the tip 209
treatment pressure 235 zipper crack 38, 40
equation 38

You might also like