Вы находитесь на странице: 1из 10

26th US Symposiumon Rock Mechanics/ Rapid City, SD / 26-28 June 1985

Energyanalysis
of hydraulic
fracturing
J.SHLYAPOBERSKY

ShellDevelopmentCompany,Houston, Texas,USA

INTRODUCTION

Hydraulic

fracturing

is

increasingly

being

used

as

stimulation

technique for oil and gas production from low permeability


reservoirs.
Most industrial
fracture
treatment designs are still
based on two simple
hydraulic
fracture models:
the Khristianovitch-Geertsma-de
Klerk (KGK)
model (Geertsma & de Klerk 1969) and Perkins-Kern-Nordgren
(PKN) model
(Perkins & Kern 1961, Nordgren 1972).
These models have been compared
and preferred
conditions for their application
to treatment design have
been discussed (Geertsma & Haafkens 1979).
Both models have two main
shortcomings.
They neglect rock strength and cannot predict
the fracture height growth.
The latter
fact has, in recent years, stimulated
intensive
research
to generate
complex hydraulic
fracture
growth and
containment models.
Computer codes are being developed by Terra Tek
(Clifton
& Abou-Sayed 1981), MIT (Cleary 1980, Cleary et al. 1983), ORU

(Palmer & Luiskutty 1984) and other research groups.

Because of simpli-

fying
assumptions about hydraulic
fracturing,
these models predict
different
fracture
geometries
for identical
reservoir
and treatment
conditions (Palmer & Luiskutty 1984).
The propped fracture
length estimated using pressure build-up
tests

and/or production data is often much less than predicted by hydraulic


fracture
simulators.
The propagating overpressure (the fracture
treatment bottom hole pressure minus the fracture closure pressure) observed
in the field
is frequently
much higher than that predicted by hydraulic
fracture simulators (Warpinski 1984, Medlin & Fitch 1983).
If a hydraulic
fracture
model is "tuned" on high propagating overpressure,
the
predicted
fracture
length will
usually
be shorter
and more consistent
with the well's
production response.
Two mechanisms could cause these high treatment pressures. One is due
to viscous flow through a much more complex multiple
fracture
system
than the planar fracture assumed in hydraulic fracture models (Warpinski
& Teufel 1984; Medlin & Fitch 1983).
Another possible mechanism is due
to a layer of relatively
small cracks around a large hydraulic fracture

that causes a significant


fracture
hydraulic

increase of the apparent fracture

growth and thus, increase


fracture
propagation.

of treatment

pressures

energy with
required

for

539

SCALE

EFFECT

IN

HYDRAULIC

FRACTURING

Scale effects
are inherent
in geomechanical
problems.
Traditionally,
mathematical
theories
describing
rock behavior
have been formulated
based on analysis
of laboratory
experiments
with small rock specimens.
Legitimate
questions
arise about the applicability
of such theories
to
predict
behavior of large rock bodies,
especially
in cases when rock
characteristics
required
by the theory demonstrate
a strong dependence
on specimen size and test conditions.
Contemporary fracture
mechanics are based on Griffith's
idea that
failure
of brittle
solids
is due to microcracks
always present in real
materials.
Griffith
formulated
a crack propagation
criterion
and introduced a new material
constant,
specific
fracture
surface
energy
,
which characterizes
material
resistance
to crack growth (fracture
toughness).
Using the Griffith
energy balance, it was shown that the critical pressure
necessary
to start
propagating
a penny-shaped crack of
radius R is (Sack 1946)

p* = /(/2)''/R

(1)

whereE' = E/(1-u2) is plainstrainYoung's


Modulus.
The energy per unit
atomic

bonds

when

the

crack area
crack

is

(
formed.

is the energy required


It

is

true

material

to rupture
constant

only for ideal


crystals.
Thus, the Griffith
theory can be, strictly
speaking,
applied
only to crack propagation
in ideal,
defect
free,
single
crystal
solids.
The applications
of the theory
to fracture
propagation
in rocks and other real
solids
are questionable.
These
solids
contain
different
types of flaws which makes the fracturing
process extremely
complicated.
Microscopic
observations
of fracture
propagation
in rocks,
polymers
and metals
show that
the material
in the region
surrounding
the main
crack undergoes a microstructural
transformation.
In rocks, an extensive array of microcracks was observed (Friedman et al. 1971); in polymers microcrazes and recrystallized
zones appeared (Bevis & Hull 1970);
in metals,
creation
of microcracks (Chudnovsky & Bessendorf 1983) and
the presence of plastic
zones, associated
with extensive
dislocation
motion, was noted.
The complicated
structure
of the damage in the material
near an
induced fracture
is typical.
Similar
features
of rock transformation
in
a zone near a propagating hydraulic
fracture
can be anticipated
only on
a large
field
scale.
Indeed,
recent
observations
in mineback experiments (Warpinski
& Teufel 1984) have shown that there are considerable
surface
roughness and wavyhess,
echelon
fracturing,
and significant
offsets
when natural
fractures
are intersected.
Detection
of hydraulic
fractures
by borehole geophones supports the presence of a relatively
narrow zone adjacent
to the fracture
with high microseismic
activity
(Hart et al. 1984).
The microseismic signals can be generated either by
shear slippage
along pre-existing
joints
or by cracks propagating
in
shear-tensile
combined mode. In any case, these signals indicate
energy
dissipation
due to irreversible
rock transformation
in a near-fracture
active
zone called
crack layer.
To properly
describe
fracture
propagation
in real solids,
theory has to be modified
to account for energy dissipated

the Griffith
in creation
of the crack layer.
It was proposed (0rowan 1952; Irwin 1958) that this
dissipated
energy be added to the specific
fracture
surface energy and a
new characteristic
called
the apparent fracture
surface energy be determined from laboratory
tests
on pre-cracked
specimens using formulas
540

similar
to (1).
This apparent
fracture
surface
energy which is an
empirical
characteristic
of fracture
toughness,
is the total
energy
required
to create a unit area of the main fracture
and the crack layer
associated
with this
area.
A difficulty
in applying
this
GriffithOrowan-Irwin
theory is that the apparent
fracture
surface
energy is not
a material
constant,
but depends on the whole history
of crack layer

development during fracture


propagation (Bakar et al. 1984).
One may
anticipate
that as the crack layer grows with fracture
propagation,
the
fracture
toughness increases
also.
This effect
of crack growth on the
rock
fracture
toughness
has been observed
in
fracture
lab
tests
(Ouchteralony
1983).
Microseismic
activity
during hydraulic
fracture
treatment,
measured by borehole
geophones and direct
observations
in
mineback experiments also demonstrate that the size of the crack layer
around hydraulic
fracture
can be much larger
than seen in lab experiments.
The corresponding
rock toughness characterizing
hydraulic
fracture propagation
should also be significantly
higher than its laboratory
measured

value.

DETERMINATION

OF FRACTURE

TOUGHNESS

FROM

TREATING

PRESSURE

This scale effect


on rock toughness in hydraulic
fracturing
process
manifests
itself
indirectly
through very high fracture
treating
pressures observed during frac jobs.
Accurate measurements of the fracture
propagation
overpressure
and comparison
with
numerically
simulated
overpressures
show that hydraulic
fracture
models, which either neglect
rock toughness or use its
laboratory
measured values,
predict
much
smaller
overpressures
than
actually
appear
in
field
conditions
(Warpinski
1984a,b).
The rock toughness characteristics
can be determined from field
pressure data by matching these and the overpressure
simulated with a hydraulic
fracture
model, in which fracture
toughness

coefficient ( K1C) is considered as a free matchingparameter. The


proposed approach requires
much effort
in field
performance and contains
uncertainties
in test procedures and data interpretation.
The fracture
closure pressure (p.),
assumed to be equal to the minimum

in situ stress, has to be determknedfirst.


An accurate estimate of
this pressure is essential for the whole analysis. To measurePc'

minifrac
and microfrac
shut-in
and flowback tests are used (Nolte 1984,
McLennan & Roegiers 1982, Warpinski
1984a).
In a shut-in
test,
the
pressure
decline
curve
has no evident
features
which would indicate
fracture
closure.
A detailed
reservoir
analysis
of the pressure falloff is thus required
to avoid a subjective
and erroneous definition
of
the closure pressure.
Different
approaches in defining
p_ from shut-in

pressure data have been reviewed (McLennon


& Roegiers 198).

An alter-

native
technique
to determine
the fracture
closure
pressure
is the
constant
rate
flowback
test.
The flowback
pressure
decline
curve
usually
has
two distinctive
points
which may indicate
fracture
closure.
The inflection
point A on the pressure decline curve (Figure
1) has been successfully
used to estimate the fracture
closure pressure

(Nolte 1984).

However, this approachmay sometimesyield unequal Pc

values in two subsequent flowback tests performed with different


flowback rates (Warpinski 1984a).
A mathematical analysis of the flow back
tests
and several
field
tests
suggests
that
point
C on the flowback
pressure decline
curve may be a better
indication
of the fracture
closure pressure because under certain
conditions
it represents
a well
reproducible
rate independent pressure value.

Once Pc is known, the fracture overpressure near the well bore,

APw

Pw Pc (Figure 2) can be calculatedfrom the bottomhole


541

pressure

(BHP) and theoretically

through perforations:

estimated

Ap
= pW - pC
W

friction

pressure

= BHP- p - pC .If

drop (p_)

the inJeg-

tion BHPdata are usedin the matchprocedure,an tificially

high rock

toughness can be expected because:


(1)
Some fluid
flow restrictions

may exist
in the perforated
zone
1984b)
which create
higher
than theoretically
estimated
friction
pressure
drop at
the
fracture
entrance
and,

(Warpinski
perforation

respectively,lowerpressure APwin the fracture.


(2)

The averaged overpressure p

]Ap(x)dA/A and the critical

stress intensity factor Ki[Ap] maybe smaller for the real fracture

than calculated
by a hydraulic
fracture
model (Figure 3).
These two effects
of fluid
flow are difficult
to quantify.
Their
significance
in the pressure analysis
can be reduced by using shut-in
pressure data.
During the early shut-in period the pressure gradient
in
the fracture
is almost eliminated
and p
= p (small pressure drop in

the fracture maystill


shut-in).

If

exist due to fluid overflowfor sometime after

one notices

that

the average overpressure

()

for

given fracture
geometry is determined by the fracture
volume and the
latter
changes insignificantly
after
a short time of pressure equalization after
shut-in,
a surprizing
result is that the ISIP is a measure of
the average
fracture
propagating
pressure,
not the fracture
closure
pressure (the minimum in situ stress).
Apparent rock toughness can be
estimated by the Griffith-Sack
type equation

KlC
where

the

fracture

= p ieff

shape dimensionless

constant

(2)

is

of

the order

of

one, the effective fracture radius Reff is the radius of an uncontained

fracture
or the half
height
of a contained
fracture.
The effective
fracture
size has to be known to use Equation (2).
In microfrac
tests,
the fracture
shape is assumed to be circular
and its
radius
can be
calculated
from pressure decline
curve (modified
Nolte analysis).
In
minifrac
tests,
the fracture
is assumed to be contained in the pay zone
wth half height considered
as the effective
fracture
radius giving a
conservative
estimate
of fracture
toughness.
FIELD

EXAMPLE

We now apply the proposed method to three minifrac


tests conducted in
the MWX Paludal
Zone Phase I stimulation
(Warpinsk 1984a).
Prior
to
these tests, the minimum in situ stress of about 40.7 MPa has been estimated by different
techniques
(microfrac
tests,
step rate
test,
flow
back test).
The pressure
decline
curve for the first
test
shows

Pc - 41.4MPaand ISIP 44.2MPa, yieldingpp- 2.8MPa.

The pay zoneheight of about 24mgives Reff= 12m,whichis less than


the temperaturesurveysshow(fracture height HT - 30min the first
test
K

1C

and H - 45m in the last one).


The use of equation (2) gives
- 9.7 MPa
or in terms of the apparent fracture surface energy
2

F = KI /E

two

othe

appearing

tests

- 3 kJ/m (E = 31GPa). Resultsof the analysisfor


summarized

in hydraulic

in

fracture

Table

treatments

indicate

that

is much larger

rock

toughness

than one would

expect to measure in laboratory fracture tests (Ouchterlony 1983); one


and two orders of magnitude for K C and F , respectively.
These

results
agree
withthegeneral
trenofheapparent
fracture
surface
energy
rowth
with c%ack
size;
.
o

Z
(

for
542

10- m),
"small"

-1 J/m for micro grain size crack

100 J/m for lab size cracks ( 10 Zm) and 10,O00 J/mZ

hydraulic

fractures

(<lOOm).

This

scale

effect

is

very

likely
to be caused by the crack layer
accompanying the hydraulic
fracture
propagation
and is clearly
seen even for small changes of
pumped volumes in Table 1.
How strong this effect can be in large
treatments
is an important practical
issue.
Some limited
data reported
in the literature
(Medlin & Fitch 1983) demonstrate tremendous increases
of post treatment
ISIP's
of more than 16MPa with
respect
to their
initial
values
which resulted
in screenouts.
It
is suggested
that
pressure screenouts be studied not only as sandout phenomena, but also
as a phenomenon of crack layer
evolution
during hydraulic
fracturing
treatments.

ENERGY

ANALYSIS

OF

HYDRAULIC

FRACTURING

The analysis
of
the
fracture
propagation
process
and hydraulic
fracturing
treatment
data suggests that the rock toughness effect
is
very important
and, under certain
conditions,
may even become dominant
for fracture
growth.
Therefore,
any hydraulic
fracture
theory has to
account for at least four interacting
processes--fluid
flow, rock deformation,
rock toughness, and fluid
loss through fracture
walls to adequately
describe
hydraulic
fracture
growth and to predict
realistic
fracture
dimensions.
An approximate
theory that incorporates
the rock
toughness effect
in conventional
fracture
models is presented below for
a circular

fracture.

A circular

fracture

equilibrium

growth.

of radius

R at time

The following

t is considered

during

quasi-

volum4ebalance accounting for the

fluid pumped
(qt), leak-offv_ume
(2RvCTr) (Nolte1984),and
:

the current fracture volume (R)

R2= qt- 2R2VCTr


Here

is

a constant (4/3

/2)dependent

on the

(3)
total

fluid

loss coefficientCT, fluid visousity, and the apparent


fracture

energy .

is

th

average fracture

width determined from

energy

considertions.

For
quasiequilibrium
fracture
propagation,
the
total
energy
dissipation
rate is minimized.
Three processes give the major contribution to the total
energy loss in the hydraulic
fracture;
(1) creation

of the new fracture

surface

and the crack layer

( .c ),

openingby deformingthe surroundingreservoir ( d )'

dissipationin fluid (f).

(2)

fracture

(3) viscous

Basedon an elasticity solutionfor a

penny-shaped crack and fluid


flow between parallel
plates,
all energy
rates can be calculated
analytically
and expressed in terms of average
fracture

width:

= rq/,d= (3/32)E'q/R,
f = (12/)t21n/()
3
(4)
c

where0 = R/Ro, 2Ro = the heightof the perforated


interval.
The condition

which yields

of quasiequilibrium

an expression

for

fracture

propagation

the average fracture

requires

width

([)2=w2+ /w+w}
(6)
where wc
2= (16/3a)m/E',
w= (1281no/(3a))qR/E'
c

543

Solving
fracture

equations (4) & (6),


the fracture
radius R and the average
width at time t are determined.
The average overpressure
Ap = (3/16)E'w/R

and the pressure

profile

for

flow between parallel

plates

p(r)= Pw-(6q/)(w--)-31n(r/R)
o

can

be

calculated.

SCALING

The

LAWS

OF HYDRAULIC

presented

FRACTURING

hydraulic

fracture

dimensionless characteristics

model

allows

introduction

with clear physica maning.

of

The .energy

balancesuggeststwo energyparametersk f = c /of andkd = f/,.

The widthAequation
suggestsa dimensioness
eometricaparameter
k = Wc/W. Finally the materialbalancegivesdimensionless
leak-

of parameter
k1 = 2RCTV_/(q/)that is relatedto the fluid effi-

ciencyef (the ratio of cretedfracturevolume


to injectedvolume)
as

k1 = 1 - ef (Nolte 1984). The three independent


dimensionless
para-

meters k

,
and k1 quantify three interactive mechanisms
betweenfour
physicalCgrockesses
involved
inhydraulic
fracturing.
Thus,
they
repre-

sent scaling laws for hydraulic


fracturing
which have to be obeyed in
properly scaled lab hydraulic
fracture
experiments.

CONCLUS ION

The scale
effect
on fracture
toughness
is discussed
in context
of
hydraulic
fracturing.
The crack layer causes this effect
and results
in
significant
increase of fracture
toughness and treatment pressure with
crack growth.
The laboratory
fracture
toughness measurements cannot

quantify rock fracture toughness of large hydraulic fractures.


A method
to estimate
fracture
toughness from treatment
pressure data is proposed
and a field
example is discussed.
A variational
principle
of the minimum energy dissipation
rate is used to incorporate
fracture
toughness in
classic hydraulic
fracture
models.
Dimensionless characteristics
of the
hydraulic
fracturing
process are introduced which are the scaling laws
of the process.

ACKNOWLEDGEMENTS

The

author

permission
discussions

wishes

to

to publish

thank

the

management

of

Shell

Oil

Company for

the paper and Dr. W. F. J. Deeg for many valuable

and assistance

in preparing

the paper.

REFERENCES

Bakar, M., A. Chudnovsky & A. Moet 1983. The Effect of Loading History
on the Fracture Toughness of Polycarbonate.
Proc. Int. Conf. on
Fatigue in Polymers. London, GB. 81.-8.8.
Biot, M.A., L. Masse & W.L. Medlin 1982. A Two-Dimensional Theory of
Fracture Propagation.
SPE Paper 11067.
57th Annual Conference.
New
Orleans, LA. September.
544

Chudnovsky, A., & M. Bessendorf 1983.


Toughness Characterization
in Steel.
Western

Reserve

Univ.

Cleveland,

Crack Layer Morphology and


NASAReport 168154. Case

OH.

Cleary, M.P. 1980.


Comprehensive Design Formula for Hydraulic
Fracturing.
SPE Paper 9259.
55th Annual Conference.
Dallas,

TX.

September.

Cleary,

M.P., M. Kawadas & K.Y. Lam 1983.

Development of a Fully

Three-Dimensional
Simulator
for Analysis and Design of Hydraulic
Fracturing.
SPE Paper 11631.
Symposium on Low Permeability
Gas
Reservoirs.
Denver, CO. March.
Clifton,
R.$. & l.S. Abou-Sayed 1981.
A Variational
Approach to the
Prediction
of the Three-Dimensional
Geometry of Hydraulic
Fractures.

SPE/DOEPaper 9879.
Denver,

Symposiumon Low Permeability

Reservoirs.

CO. May.

Friedman,

M., J. Handin & G. Alani

1972.

Fracture-Surface

Energy of

Rocks.
Int.
J. Rock, Mech. 9:757-766
Geertsma, J. & F. deKlerk 1969.
A Rapid Method of Predicting
Extent of Hydraulically
Induced Fracture.
JPT December.

Geertsma, J. & R. Haafkens 1979.


Predicting

Width

Fractures.

Trans.

and Extent
ASME.

J.

A Comparison of the Theories

of Vertical
En.

Width and

Res.

Hydraulically
Tech.

101:8-19.

for

Induced
March.

Hart, C.M., D. Engl, R.P. Fleming & H.E. Morris 1984. Fracture
Diasnostics Results for the Multiwell
Experiments's Paludal Zone
Stimulation.
SPE/DOE/GRI Paper 12852. Unconventional Gas Recovery
Symposium.
Pittsburg,
PA. May.
Irwin,
G.R. 1958.
Analysis of Stresses
Crack Traversing
a Plate,
Discussion.

and Strains Near the End of a


Trans. ASME. J. Art. Mech:

299-301.

Mclennan, J.D.
Instantaneous
Conference.
Medlin, W.L. &
Treatments.
CA.

& J.C. Roegiers 1982.


How Instantaneous
are
Shut-In Pressures.
SPE Paper 11064.
57th Annual
New Orleans, LA. September.
J.L. Fitch 1983.
Abnormal Treating Pressures in MHF
SPE Paper 12108.
58th Annual Conference.
San Francisco,

October.

Nolte, K.G. 1984.


A General Analysis of Fracturing
Pressure Decline.
SPE Paper 12941.
Nordgren, R.P. 1972.
Propagation of a Vertical
Hydraulic Fracture.
SEPJ:

306-314.

Orowan, E. 1952.
Fundamentals of Brittle
Behavior in Metals.
In W. M.
Murray (ed.),
Fatigue and Fracture of Metals.
New York:
John Wiley &
Sons.

Ouchterlony,
F. 1983.
Fracture
Toughness Testing of Rock. In H. P.
Rossmanith (ed.),
Rock Fracture Mechanics.
Int.
Centre Mech. Sci.
CISM Courses & Lectures No. 275. Wien - New York:
Springer.
Palmer, I.D. & C.T. Luiskutty 1984.
Comparison of Hydraulic Fracture
Models for Highly Elongated Fractures
of Variable
Height.
Oral
Roberts Univ. Preprint.
Tulsa, OK.
Perkins,
T.K. & L.R. Kern 1961.
Width of Hydraulic
Fractures.
JPT: 937949

Warpinski,
N.R. 1984a.
Summary of Results of MWX Paludal Zone Phase I
Stimulation.
Memorandumof Record.
Sandia National Labs. January.
Warpinski,
N.R. 1984b.
MWX Paludal In Situ Stress Measurements and
Hydraulic
Fracture
Behavior.
Memorandum. Sandia National
Labs.
October.

Warpinski, N.R. & L.W. Teufel 1984. Influence of Geological


Discountinuities
on Hydraulic Fracture Propagation.
SPE Paper
13224.
59th Annual Conference.
Houston, TX. September.
545

Table 1.
Paludal

Apparent Fracture
Zone

Phase

Toughness From Mini Frac Tests in WX

Stimulation*

Volume
ISIP
(m3)
(MPa)

Ap
(MPa)

(HTm
Ref
K1C
)
(m)f
MPayr

22

44.2

2.8

28

12

9.7

57

45.9

4.5

41

12

15.6

47.6

6.2

46

12

21.5

114

KJ/m
2
3

7.8
14.9

*Pc 41.4 MPa,E' = 31 GPa,HpA


Y = 24m(Warpinski1984a)

qB

"1'

INJECTION

BHP pwJ5
PT

HUT - )N

ISlp
-p(

BHP
,,=DW
' PT
FLOW

BACK

ir

A B

'///////

qB'

Figure 1.
Injection
- shut-in
flowback pressure curve of typical
micro/mini
fracture
test.

Figure 2.
Bottom hole pressure
measurements and friction
drop
during fracture
test.

IP-

DISTANCE

ALONG THE FRACTURE

DISTANCE

ALONG

THE FRACTURE

Figure 3.
Hypothetical
pressure profile
during injection
for (A) idea1 well-fracture
system, (B) fracture treatment.
546

and shut-in

13. Poster session A

Вам также может понравиться